Next Article in Journal
Anti-Inflammatory Effect of Cudrania tricuspidata Extract and Stewartia koreana Extract Mixture in a Collagen-Induced Arthritis Mouse Model
Next Article in Special Issue
Assessment of Human Pharmaceuticals in Drinking Water Catchments, Tap and Drinking Fountain Waters
Previous Article in Journal
Beetroot Microencapsulation with Pea Protein Using Spray Drying: Physicochemical, Structural and Functional Properties
Previous Article in Special Issue
Notes on Common Misconceptions in Microplastics Removal from Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Removal of Pharmaceuticals from Water by Adsorption and Advanced Oxidation Processes: State of the Art and Trends

1
Laboratory of Environmental Engineering and Ecotechnology, National School of Engineers (ENIS), Sfax University, Route de la Soukra km 4, Po. Box n° 1175, 3038 Sfax, Tunisia
2
Higher Institute of the Sciences and Techniques of Waters, Gabes University, 6072 Zrig Gabes, Tunisia
3
Aix-Marseille University, CNRS, IRD, INRAE, Coll France, CEREGE, CEDEX, 13454 Aix-en-Provence, France
4
International Water Research Institute, Mohammed VI Polytechnic University, 43150 Benguerir, Morocco
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(14), 6659; https://doi.org/10.3390/app11146659
Submission received: 15 June 2021 / Revised: 14 July 2021 / Accepted: 17 July 2021 / Published: 20 July 2021

Abstract

:
Pharmaceutical products have become a necessary part of life. Several studies have demonstrated that indirect exposure of humans to pharmaceuticals through the water could cause negative effects. Raw sewage and wastewater effluents are the major sources of pharmaceuticals found in surface waters and drinking water. Therefore, it is important to consider and characterize the efficiency of pharmaceutical removal during wastewater and drinking-water treatment processes. Various treatment options have been investigated for the removal/reduction of drugs (e.g., antibiotics, NSAIDs, analgesics) using conventional or biological treatments, such as activated sludge processes or bio-filtration, respectively. The efficiency of these processes ranges from 20–90%. Comparatively, advanced wastewater treatment processes, such as reverse osmosis, ozonation and advanced oxidation technologies, can achieve higher removal rates for drugs. Pharmaceuticals and their metabolites undergo natural attenuation by adsorption and solar oxidation. Therefore, pharmaceuticals in water sources even at trace concentrations would have undergone removal through biological processes and, if applicable, combined adsorption and photocatalytic degradation wastewater treatment processes. This review provides an overview of the conventional and advanced technologies for the removal of pharmaceutical compounds from water sources. It also sheds light on the key points behind adsorption and photocatalysis.

1. Introduction

In the past three decades, pharmaceutical residues have been discovered in almost all environmental matrices on every continent [1]. Approximately 3000 pharmaceutical substances are used in the European Union. Pharmaceuticals have been detected in groundwater, urban wastewater, surface water, and drinking water in a range of ng to μg per litre [2,3,4,5,6]. Because of their low biodegradability and high hydrophilicity, pharmaceuticals are difficult to eliminate from water systems using conventional wastewater treatment techniques [7]. Therefore, their disposal in wastewater treatment plants is a major concern [6,8]. Indeed, pharmaceuticals are considered the most significant groups of environmental pollutants of special concern. The most consumed type of pharmaceutical products may differ from one country to another. For example, in Spain, antihistamines, analgesics, and antidepressants are the most consumed based on the National Health System [9]. However, in Italy, antibiotics, sulphamethoxazole, ofloxacin and ciprofloxacin, β-blocker atenolol, antihistaminic ranitidine, diuretics furosemide, hydrochlorothiazide, and steroidal anti-inflammatory drugs (NSAIDs) such as ibuprofen are the most detected pharmaceuticals in sewage treatment plants [10]. Klein et al. [11] reported that Tunisia was considered the second-highest consumer of antibiotics in 76 countries analyzed between the years 2000–2015, and consumption was estimated at around 47 defined daily doses per 1.000 inhabitants per day.
Many studies have shown that wastewater treatment plants (WWTPs) do not totally remove pharmaceutical compounds [9]. More than 100 different drugs have been found in the aquatic environment [3,12]. Antibiotics and NSAIDs were reported in surface water samples from various countries at concentrations ranging from 5 to 150 µg/L [9,13,14]. Castiglioni et al. [10] noted that there is a difference in concentrations in pharmaceutical compounds in WWTPs between winter and summer. This is mainly due to greater attenuation and lower use of pharmaceuticals in summer. Thus, pharmaceuticals and their metabolites are detected regularly in aquatic and terrestrial environments [15,16]. Hence, tertiary treatment is needed to eliminate these emerging pollutants. Today, countries are facing high COVID-19 pandemic incidence rates and struggling to manage the dramatic increase in medical waste production by healthcare facilities, in particular with respect to pharmaceutical products. For instance, Wuhan inhabitants in China (~11 M) produced ten times more daily medical waste than the previous average (200 tons on a single day, 24 February 2020). A drastic increase in medical waste was also reported in other parts of the world, such as in Catalonia, Spain, and in China, with increments of 350% and 370%, respectively [17]. Recent studies indicated that the COVID-19 pandemic has led to an increase in waste generation by an average of 102.2 % in both private and public hospitals [18]. In addition, the hazardous waste volume in the hospitals’ investigations has increased by an average of 9% in the amount of medical waste and by 121% compared to the first wave and before the epidemic COVID-19 [19,20].
The hydrophobic/hydropholic nature of pharmaceutical compounds facilitates their interaction with microplastics, frequently present in surface water and playing the role of a vector of pharmaceuticals within aqueous environments through π–π interactions [21]. The literature cites about 160 pharmaceutical compounds that have been recently found in the water surface and wastewater [21]. There are already approaches and strategies for minimizing this issue that proved their efficiencies and estimates to be implemented and prioritized in the near future. Currently, the accuracy of techniques for pharmaceutical removal from wastewater differs from drugs to other (80–100%). With this purpose, the development of new technology such as combined techniques based on an in situ census is still needed.
The main focus of this review is an in-depth compilation and discussion of the treatment options investigated for the removal of pharmaceuticals, including methods of detection and removal of pharmaceutical products from water sources. To this point, the information gathered in this review provides a clear concept in a single report to enable a fast understanding of the current status of the pharmaceutical detection field. This paper summarizes the methodological, socio-economic, and technological factors in both the short and long term influencing the pharmaceutical loading and removal from wastewater. A critical discussion of these approaches for implementation in wastewater treatment processes is also presented. So far, there are still major gaps between laboratory and field-scale research that need to be overcome in order to assess the viability of real applications.
The ultimate goal of this review is to provide a platform on the recent development in the coupling of adsorption and advanced oxidation processes for the removal of PPCPs from the aquatic environment. The pharmaceutical adsorption capacity by adsorption and advanced oxidation process and their combination is discussed, and their efficiencies on the concentration of pharmaceuticals removed are described in this review. Continuing efforts in the following areas will strengthen the potential of the coupled process AOPs/adsorption in PPCPs removal from source water.

2. Commonly Used Pharmaceuticals: Their Classes, Chemical Structures, and Therapeutic Applications

Approximately 3000 pharmaceutical substances are used in the European Union. The most widely used molecules are antibiotics for human and veterinary medicine, their consumption has reached 12,500 tons per year over the last decade. Non-steroidal anti-inflammatory drugs (NSAIDs) such as acetylsalicylic acid (aspirin), diclofenac, and ibuprofen are inhibitors of the two isoforms of cyclo-oxygenase (cox-1 and cox-2) enzymes, involved in prostaglandin synthesis. Ibuprofen, scientifically named 4-isobutylphenyl-2- propionic acid, is one of the most widely used NSAIDs worldwide. Additionally, there are other types of drugs, such as analgesics like paracetamol, that are considered antipyretic (Table 1). It appears to inhibit cyclo-oxygenase in the central nervous system without peripheral action, which is the reason why it does not have anti-inflammatory effects. Naproxen (NPX) is one of the most effective analgesics, a non-steroidal anti-inflammatory drug (NSAID), and it has been found in both surface water and wastewater at concentrations ranging from 20 ng/L to several mg/L [22]. Additionally, it is detected in wastewater treatment plants in the range of 250 to 1.5 μg/L, and its removal has been estimated at around 71% [23]. NPX has an excellent cardiac profile; however, it has been reported that people who ingest trace amounts of NPX have a higher risk of heart attack or a higher potential of urinary bladder cancer [24,25,26]. Depending on the development of their health services, the type and the concentration of drugs in sewage systems vary in each country. As an example, in Table 1, major pharmaceutical products used in Tunisia are shown with a predominance of antibiotics and analgesics. In Northern countries, synthetic and natural hormones, such as 17-α-ethinylestradiol (EE2) and 17-β-estradiol (E2), are present in wastewater at a non-negligible level [27].

3. Concentration of Pharmaceuticals in Water Environment

Numerous pharmaceutically active compounds have been discovered in water since the first discovery of such contaminants in aquatic systems in the 1980s [1]. Human drug residues occur mainly in hospital effluents and, in the case of urban medicine, in sewage systems containing patient discharge. These effluents loaded with residues are treated in the treatment plant before reaching the aquatic environment. Residual pharmaceuticals may introduce toxicity into our population, ecosystems, or the ecosphere [29]. The presence of pharmaceutical compounds in the environment, in the form of non-modified active ingredients, has been known since the first studies on a few compounds in the effluents of wastewater treatment plants or aquatic environments [30]. They can reach surface water, groundwater, and then drinking water [9].
In general, NSAIDs are the most consumed all over the world because of their availability without medical prescription [9]. For example, ibuprofen is still detected at μg/L concentrations [31,32,33,34]. Despite the fact that the removal rate of this medicine is very high, it is always detected in the river downstream. Likewise, two additional compounds have been detected: acetaminophen (paracetamol) and aspirin (acetyl-salicylic acid) [9]. Akhtar et al. [35] indicated that only 20% to 30% of pharmaceuticals for anti-inflammatory, β-blockers, and analgesics could be removed, which was in contrast to antibiotics, where 50% can be eliminated from the effluent water.
According to Moslah et al. [27], in Tunisia, antibiotic and β-blockers are the most observed groups in wastewater, which means that these compounds are resistant to biological WWTPs. For antibiotics, their concentrations in the effluent ranged from 35 ng/L to 1.2 µg/L. Moslah et al. [27] reported that the pharmaceutical products that represent the most consumed drugs in Tunisia are paracetamol, mefenamic acid, and ibuprofen. Moslah et al. [27] also reported that other pharmaceuticals such as antibiotics and β-blockers were detected in the effluents of WWTPs, and their concentration ranged from 100 to 1 μg/L. Moreover, they found that the concentration levels of erythromycin, ofloxacin, and sulfamethoxazole in effluent samples were greater than those in influent samples, which can be explained by the deconjugation processes that may occur during their contact with activated sludge [36,37].
The interaction between microplastics (MPs) and pharmaceutical residues is an issue of increasing concern in various environmental compartments. Indeed, MPs can change the effect of pharmaceutical distribution and their transport through the sorbent and sorbate pair [38]. Microplastics coated with pharmaceutical residues have been transported across marine, freshwater, and terrestrial ecosystems [39]. The MPs with these contaminants pollute pristine ecosystems, leading to ingestion by marine and aquatic organisms, which can be lethal to marine life [40]. Data on the adsorption of pharmaceutical residues on the MPs surface are quintessential for risk and environmental assessment. Several studies showed that the desorption rates of polyethylene MPs increased in the order of propranol, sulfamethoxazole, and sertraline, which were associated with the increment of the adsorption capacity and the hydrophobicity [21]. In addition, Lin et al. [41] revealed that marine and warm-blooded organisms are at higher toxicity exposure to atorvastatin, amlodipine, and tetracycline even at a lower concentration due to their chronic degradation and metabolites, which is in contrast to organisms that are cold-blooded.
In general, the reuse of wastewater (WW) for the irrigation of vegetable crops is widely applied all over the world [42]. In addition, pharmaceuticals can be obtained through the consumption of plants, vegetables, fruit, fish, and meat [42]. Some studies carried out by Christou et al. [42] show that different concentrations of pharmaceutical residues exist in plants (tomato) successively by sulfamethoxazole SMX (0.98 mg/kg) followed by trimethoprim TMP (0.62 mg/kg) and diclofenac DIC (0.35 mg/kg). However, in the first year of the study, DIC was not strongly detected. However, DIC showed the highest concentration (11.63 mg/kg) throughout the study, followed by SMX (5.26 mg/kg) and TMP (3.40 mg/kg). Very high concentrations of ciprofloxacin (31 mg/L), cetirizine (1.3–1.4 mg/L), and losartan (2.4–2.5 mg/L) were reported in the effluent of a treatment plant serving 90 bulk drug manufacturers in Patancheru, near Hyderabad, India. Enoxacin (150–300 μg/L), Citalopram (770–840 μg/L), lomefloxacin (150–300 μg/L), enrofloxacin (780–900 μg/L), metoprolol (800–950 μg/L), ofloxacin (150–160 μg/L), norfloxacin (390–420 μg/L), and ranitidine (90–160 μg/L) were also found in this study. Clearly, this wastewater effluent is of great concern to local microorganisms and humans, where they are directly and indirectly exposed. The total release of ciprofloxacin (at 31 mg/L) into the environment is 44 kg per day. This release is equivalent to the total consumption of Sweden over five days (population of nine million). The evolution of drug-resistant bacteria is favored in these waters.
In fact, when pharmaceutical residues are thrown into the environment from untreated wastewater containing various toxic drugs, people will be directly exposed via the consumption of plants that are already irrigated or via the consumption of contaminated fish (Figure 1). As long as we have not found solutions or undergone advanced treatments for the elimination of dangerous pharmaceutical residues, man will always remain a victim of his own actions.
Khan et al. [43] studied the abundance of antibiotics in the North of Pakistan in 19 sampling sites (river, dam, canal, sewage drain, etc.). They found high levels of antibiotics in a river in the close vicinity and downstream of Lahore city (10 million inhabitants). The highest concentration of antibiotics for foroxytetracycline, trimethoprim, and sulfamethoxazole was equal to 1100, 1700, and 2700 ng/L, respectively. They noted that the presence of antibiotics in surface water can be horizontally transferred to human-associated bacteria and thus contribute to antibiotic resistance proliferation.

4. Removal of Pharmaceuticals during Wastewater Treatment

4.1. Conventional Treatments

Conventional wastewater treatment consists of a combination of physical, chemical, and biological processes and operations to remove solids, organic matter, and, sometimes, nutrients from wastewater. General terms are used to describe different degrees of treatment in order of increasing treatment level; preliminary, primary, secondary, and tertiary and/or advanced wastewater treatment. In some countries, disinfection to remove pathogens sometimes follows the last treatment step [44]. It has been shown that pharmaceuticals can be removed only partially, as these wastewater treatment plants are not designed for full removal [1]. Depending on the composition of wastewaters, preliminary treatments with activated sludge systems or classic membrane bioreactors may be needed [45]. However, membrane fouling comprises a significant obstacle to their broad application [46]. Pharmaceutical compounds are characterized by their limits of elimination by volatilization because of their low vapor pressure and pKa values between 3 and 10. Some drugs, such as ibuprofen, diclofenac, and carbamazepine, include extremities vulnerable to biodegradation and sorption. According to the review of Petrie et al. [47], diclofenac is removed by ≤50%, and any carbamazepine removal is low. From the study carried out by Tauxe-Wuersch et al. [48], in Switzerland, they noted the difficulties faced in the removal of different drugs such as ibuprofen, mefenamic acid, and diclofenac with biological and physico-chemical treatments. Additionally, this work reported that there was a difference in the elimination rates within wet and dry seasons. The removal of ibuprofen and ketoprofen was inhibited during winter in comparison to that during the dry period. This can be explained by the difference in the residence time of treated water in treatment plants depending on the rainfall. In fact, the flow rate was three times higher in winter than in the dry period. In addition, Lindqvist et al. [49] found ibuprofen, naproxen, ketoprofen, diclofenac, and bezafibratein in seven different sewage water treatment plants in Finland. Despite their effort to remove these pharmaceutical products, they detected them again in the rivers of discharge of sewage water treatment. In fact, in the river of Aurajoki (AUJ), they found low concentrations of all studied pharmaceuticals, but in the river of Seinajoki (SEJ, tributary of Kyronjoki, KYJ), they found a very low concentration, which was explained by the dilution of water up to five times. In the Kokemäenjoki river (KOJ), located in the South Westo f Finland, they found only IBU, NPX, and DIC, which was explained by dilution and also degradation and adsorption processes. The highest removal rate was attributed to ibuprofen with 92% and the lowest one to diclofenac with 26%. In the same context, Metcalfe et al. [50] found analgesic/anti-inflammatory drugs such as ibuprofen and naproxen, as well as the metabolite of acetylsalicyclic acid, salicylic acid, in final effluents at µg/L levels from 14 sewage treatment plants from the provinces of Manitoba, British Columbia, Quebec, Alberta, Ontario in Canada. The most important parameter for drug degradation is the sludge retention time (sludge age). When the sludge age is 10 to 15 days, partial biodegradation is observed, but after only four days, biodegradation of pharmaceuticals is almost zero [51].The sludge age required for degradation of aspirin, ibuprofen, bezafibrate, and sulfamethoxazole is 2 to 5 days [52], although, for iopromide, roxithromycin, and diclofenac, 5 to 15 days are required for significant degradation [53]. Some products such as carbamazepine and diazepam remain unmodified even after >20 days [54]. Hydraulic retention times >27 days and sludge retention times >35 days have led to high removal of estrogens from water. From the results of the above-cited studies, it can be concluded that conventional WWTP settings, which have been performed to reduce the contamination of water, did not lead to a desirable quality of water. Indeed, pharmaceutical residues were not totally removed by these processes, and they are increasingly dangerous for humans and the environment. It is of interest to explore the technological and scientific gaps in wastewater treatment by the conventional techniques with the specific study. The review paper published by Grandclement et al. [55] reported that micropollutant removal appears to be compound- and process-dependent in all investigated processes. This paper also reported that even though some hybrid processes show promising micropollutant removal, further studies are needed to optimize these water treatment processes, in particular in terms of technical and economic competitiveness.
Conventional drinking water plants feature multiphase treatments, including clarification, filtration, and coagulation/flocculation steps [1], Not one of these operations have shown an efficient removal of pharmaceutical products among the sixteen different medicines considered in this review [56]. When considering sewage water, Kooijman et al. [56] demonstrated the ineffectiveness of the coagulation–flocculation process for the elimination of pharmaceutical products. However, the use of iron (III) chloride as a flocculant product showed no significant removal for diclofenac, carbamazepine, bezafibrate, and clofibric acid from water [53]. Stackelberg et al. [57] observed low removal of sulfamethoxazole and acetaminophen up to 25%. According to Choi et al. [58], coagulation was shown to be effective for the removal of seven tetracycline classes of antibiotics (TAs), and the addition of poly-aluminium chloride (PACl) allowed authors to remove 43~94% of TAs at the optimum conditions from the synthetic water. Overall conventional water and wastewater treatment processes are unable to reliably remove some recalcitrant pharmaceuticals, and it is necessary to introduce additional advanced treatment technologies prior to discharge into the environment.
Among the emerging treatment options, combined adsorption and AOPs are the most promising methods for the efficient removal of pharmaceutical residues. The combined adsorption–-AOPs systems are discussed, which hold great promise to provide international alternatives for the manifold WWT process to mitigate pharmaceutical residues and meet acceptable limitations.

4.2. Biological Removal of the Pharmaceuticals

Bacteria occupy a dominant place in the biosphere; through their various metabolic capabilities, they support the metabolic cycles that are fundamental to all life on Earth [59]. Researchers are investigating the potential of different bacteria to degrade pharmaceuticals and personal care products (PPCPs) into eco-friendly monomers, which could be an alternative emerging way to eliminate pharmaceutical residues from the ecosystem [60]. Mainly related to the bacterial populations concerned, technologies that use sludge (aerobic, activated, or granular) have been highly efficient in the treatment of municipal wastewater [61]. A study evaluating the efficiency removal of pharmaceutical products such as atorvastatin, caffeine, paracetamol, xylazine, trimethoprim, sulfamethoxazole, naproxen, fluoxetine, diclofenac, ibuprofen, clarithromycin, carbamazepine, atenolol, azithromycin, erythromycin, ketoprofen, metoprolol, erythromycin, ciprofloxacin, valsartan, simvastan, and losartan, using activated sludge in the laboratory, at concentrations ranging from 13.2 ng/L to 51.8 μg/L in Latvia in municipal wastewater highlighted that biostimulation through the addition of nutrients to the sludge showed biodegradation of these pharmaceutical contaminants by Nitrospirae, Actinobacteria, Verrucomicrobia, Firmicutes, Acidobacteria, Proteobacteria, Chloroflexi, and Bacteroidetes. As well, some genera and species involved in removal processes of pharmaceutical compounds such as metoprolol, venlafaxine, fluoxetine, alprenolol, propranolol, bisoprolol, salbutamol, norfluoxetine, gemfibrozil, and 17β-estradiol have demonstrated a 90% or higher removal efficiency [62]. Another study improved the efficiency of using a Gram-positive bacterium, designated as strain B1 (2015b) for the degradation of ibuprofen and naproxen, after six days for 20 mg/L of concentration and 35 days for 6 mg/L, respectively [63].
The biodegradation of PPCPs could be a potential solution to eradicate this residue. This process is carried out by micro-organisms, which are more likely to decompose this pharmaceutical waste into biomass, methane, carbon dioxide, water, and various inorganic compounds, while the enzymes of organisms play a vital role. Pharmaceutical removal by biodegradation depends on what configuration is needed. The reactor configurations can be as simple as a sand column, WWTPs, and constructed wetlands, sequencing batch reactors, membrane bioreactors, which have been studied by many researchers [64]. Environmental parameters, including temperature, sunlight, atmospheric humidity, and ultraviolet rays, have also been shown to influence biodegradation [65,66,67].

4.3. Electrocoagulation Process

Electrocoagulation is a treatment process in which cations are formed by metal electrodes in an electrical field [68,69]. It produces coagulants using metal electrodes that are more likely to encounter pharmaceutical products with several advantages, including sludge minimization, automatic treatment, and efficient and low operating costs [69,70]. The removal efficiency of the three most consumed pharmaceutical products such as carbamezapine (70%), diclofenac (90%), and amoxicillin (77%) was investigated by Ensano et al. [71] for a density of 0.5 mA/cm2, an initial concentration of 10 mg/L, and a hydraulic retention time HRT of 38 h. Oxytetracycline, an antibiotic drug, was analyzed by Nariyan et al. [72]. The optimum current density was 20 mA/cm2 for both anodes: iron and aluminum had a removal efficiency of 93.2% and 87.75%, respectively. The initial concentration on removal efficiency was also studied: increasing the initial concentration of oxytetracycline hydrochloride up to 200 mg/L did not have a significant impact on its removal. The pH, Eh, and dissolved oxygen of all samples were measured during the experiments: with both anode–cathode combinations, pH was seen to increase considerably, while Eh and dissolved oxygen decreased substantially. Furthermore, the total removal of ciprofloxacin using electrocoagulation process was achieved at a density of 15 mA.cm−2, pH 7.5, initial CIP concentration of 60 mg.L−1, electrolyte dose of 0.07 M NaCl, and inter-electrode distance 1.58 cm within the equilibrium time of 20 min [71,73].

4.4. Sorption Process for the Removal of Pharmaceutical Products

The adsorption process by solid adsorbents shows potential as one of the most efficient for the treatment of a wide range of waters and wastewaters [1] that contain pharmaceutical products. In fact, there are excellent studies of carbon materials used to remove pharmaceutical pollutants, which is a financially attractive alternative for wastewater treatment.
Different types of adsorbents are classified into natural and synthetic adsorbents. Natural adsorbents include charcoal, clays, clay minerals, zeolites, and ores. These natural materials, in many cases, are relatively low-cost and abundantly available and have significant potential to modify and ultimately improve their adsorption capacities. Synthetic adsorbents are adsorbents prepared from agricultural products and waste, household waste, industrial waste, sewage sludge, and polymeric adsorbents. Each adsorbent has its own characteristics such as porosity, pore structure and the nature of its adsorbent surfaces. Many wastes used include fruit waste, coconuts, date nuts, used tires, bark and other tannin-rich materials, sawdust, rice husk, petroleum waste, fertilizer waste, fly ash, sugar industry waste, blast furnace slag, chitosan and seafood processing waste, and algae, peat, clays, red mud, zeolites, sediment and soil, ores, etc. [74].
The current application of advanced materials for the removal of PPCPs from wastewater can be viewed from three main perspectives: activated carbon (AC), clays. and biochar. Other more recent materials have also been reported, such as molecular imprinted polymers (MIPs), thermo-responsive gel, and magnetic nanoparticles, though comparatively, they still lack development and supporting case studies.

4.4.1. Activated Carbon

Removal of naproxen has been determined by different adsorbent materials and has shown different results. A compilation of pharmaceutical residue counts and removal efficiencies from literature is provided in Table 2. The more efficient adsorbent used was waste apricot plants from Malatya as activated carbon using a batch adsorption process. In fact, almost total elimination was found when the temperature increased to 50 °C for 100 mg/L [75]. In addition, the use of green synthesis of silver-reduced graphene (Ag-RGO) showed a good performance for the removal of naproxen NPX for 92.62% [76] for a quantity of adsorbent of 20.2 mg at pH of 4.5 and temperature of 24.25 °C. At the same condition of pH and temperature, the green synthesis of copper nanoparticles (CuNPs), as adsorbent materials, indicated an elimination of 86.9% after 60 min of experience for 10 mg of CuNPs. [77].
As well, activated carbon cloths can be exploited for the removal of ibuprofen according to Guedidi et al. [78]. This adsorbent was provided by Kuraray Chemical Co. Ltd. (Japan). They analyzed the removal of these drugs by distilled water and observed that the efficiency of elimination depends on the decrease of pH.
Since the research carried out for diclofenac DIC drugs, more effective adsorption has been achieved using an olive waste cake as activated carbon [79]. For a PH = 4.5, Husein et al. [77] noted that a 91.4% of removal in 60 min by using green-synthesized copper nanoparticles (Cu NPs) as the adsorbent for a maximum capacity was 36 mg/g. The smallest percentage of elimination noted for a pH = 2, dose of Cyclamen persicum tubers CTAC = 0.7 g and a concentration of diclofenac 70 mg/L was 81% for removal after 120 min of contact time [2]. For antibiotic drugs, amoxicillin was totally removed from hospital wastewater and aqueous solutions by using magnetic adsorption prepared from olive kernel MA-OK situated in the southern regions of Iran produced from palm [80] for a maximum concentration of adsorbent and AMX 0.5 g/L, 200 mg/L, respectively, at a contact time of 90 min and pH of 6 [80]. For a better understanding of the in-depth mechanistics of the removal of pharmaceutical compounds by adsorption, Mansouri et al. [8] reported that there is no difference in the adsorption sites and kinetics for IBP and amoxicillin (AMX), but IBU is more adsorbed than AMX whatever the nature of the adsorbent. Authors have demonstrated that there is a relationship between the removal of IBP and the adsorption of ionic species present in the wastewater as opposed to AMX. Hasan et al. [81] studied the adsorption of two typical PPCPs such as naproxen and clofibric acid onto the metal–organic framework (MOFs), functionalized with aminomethanesulfonic acid (AMSA) and ethylenediamine (ED) as acidic and basic groups, respectively. They concluded that the highest efficiency for adsorption rate and adsorption capacity was obtained with basic ED-(MOFs. They explained such behavior by an acid–base interaction between the PPCPs and the adsorbents. They showed also that the effect of pH on the adsorption of naproxen supports this acid–base interaction.

4.4.2. Clays

However, clay minerals could be inexpensive and widely available materials for the removal of pharmaceuticals [82]. For clay materials, the most common types (illite, kaolinite, vermiculite, montmo-rillonite, bentonite, and sepiolite) have already been used for the removal of a variety of organic micro-pollutants. Montmorillonite is the most promising adsorbent for further investigations aimed at testing the practicability of a clay-based adsorbent for the removal of pharmaceuticals compounds [82]. The use of natural clay (montmorillonite (Mt), vermiculite (VER), bentonite (B), kaolinite (K)) and modified clay-based sorbents commercial acid-activated montmorillonites K10 and K30, and two carbonaceous-mineral nanocomposites (MtG5%T, BAlG3%C) for the removal of ibuprofen, diclofenac, ketoprofen, carbamazepine, bisphenol A, and triclosan, showed that vermiculite was the best adsorbent for the removal of all drugs. For the modified clay minerals, the best results were achieved for carbonaceous bentonite, and the two acids activated montmorillonites K10 and K30 [83]. Furthermore, the adsorption of gemfibrozil, mefenamic acid, and NPX using two different clays materials, exfoliated vermiculite and LECA, showed that both adsorbents were able to remove drugs from the aqueous medium, while vermiculite presented a high adsorption capacity up to 70% [84].

4.4.3. Biochar

Biochar has been used as an adsorbent of pharmaceutical compounds by Patel et al. [1]. They demonstrated that many factors affect pharmaceutical adsorption on biochar that are common to most sorbents: pH, temperature, feedstock, preparation conditions, presence of ions, humic acid, pharmaceutical structures, and functional groups present both in the biochar’s surface and within its solid structure. Biochar-based materials have been used by Peiris et al. [85] for the removal of antibiotic sulfonamides and tetracyclines in aquatic environments. They showed that in contrast to active carbon (AC), biochar is considered a successful adsorbent; even with low surface area, it can have excellent potential as an adsorbent due to its ability to swell in water leaving more room for adsorbents [86]. Therefore, the adsorption of pharmaceuticals onto AC often results in both high removal efficiencies and capacities without generating toxic products. Hydrophobic pharmaceuticals tend to reach equilibrium with AC faster than hydrophilic compounds [1].
Table 2. Removal of pharmaceutical products by adsorption process.
Table 2. Removal of pharmaceutical products by adsorption process.
Operation ProcessDrugsAdsorption ProcessScaleMatrixExperience Time (min)Reactor Volume (mL)pHT(°C)BET (m2/g)Capacity (mg/g)Significant FindingsReferences
Batch adsorptionIbuprofen
Naproxen
Diclofenac
Green-synthesized
copper nanoparticles (Cu NPs)
Lab-scaleWastewater6010004,524,85 33.9
33.9
36
The percentage of removal of IBU for about 74.4%, for DIC about 91.4% and NPX about 86.9%.[77]
NaproxenAg-RGO Nano-composite filmLab-scaleAqueous solution3504.524,85 229.2592.62% of removal.[76]
AC from gooseberry seed-shellsLab-scaleWater treatment plants15504.40 1328154.9861.99% of removal.[87]
GB-GP-AgNs645
Ibuprofen
Amoxicillin
OP: Physical activation of the carbonized precursor using CO2Lab-scaleAqueous solution601009.5
25105586.2IBU is more adsorbed than AMX.[8]
OC: Chemical
activation of the raw precursor using phosphoric acid
5.6110678.8
OPox: Oxidation
of OP in a saturated solution of ammonium persulfate in 4 N sulfuric acid
3.490375.4
DiclofenacCyclamen persicum tubers activated carbon
(CTAC)
Lab-scaleAqueous solution12050615Zinc chloride activated: 880.936Zinc chloride activated: 606.78Elimination of 81% when DIC concentration was 70 mg/L and 0.7 g CTAC.[2]
Physical activation: 799.028Physical activation: 522.07
Gemfibrozil
Mefenamic acid
Naproxen
Exfoliated vermiculiteLab-scaleAqueous solution12050007.4720 LECA have a higher absolute removal of drugs then exfoliated vermiculite.[84]
LECA9.04
Naproxen Clofidric acidAminomethanesulfonic acid (AMSA- MIL-101)Lab-scaleDeionized water1.1050<4
>6
2523223764A higher rate constant for adsorption (k2) was found for ED-MIL-101 compared to the virgin and AMSA-MIL-101.[81]
ethylenediamine (ED- MIL-101)1.1725554159
Salicylic acid IbuprofenPine wood biocharLab-scaleAqueous solution 10002.5
3
45
35
1.3510.74Methanol was able to achieve 93% and 88% desorption of salicylic acid and ibuprofen.[88]
NaproxenWaste apricotLab-scaleAqueous solution6010005.82251060106.38For the removal of naproxen, the Langmiur equation that best represents the equilibrium data was found.[75]
Column adsorptionIbuprofen
Diclofenac
Naproxen
Ketoprofen
Olive-waste cakeLab-scaleDeionized water19.13
17.00
12.07
11.20
3004.12579310.83
56.17
39.52
24.69
90.45% of removal for NPX
88.40% of removal for ketoprofen
70.07% of removal for IBU.
[79]
IbuprofenActivated carbon clothLab-scaleDistilled water60012.57251910491.9The adsorption kinetics of IBP on the raw carbon cloth were found to be accelerated by the decrease in pH.[78]

4.5. Removal of Different Pharmaceuticals by Advanced Oxidation Processes

Klavarioti et al. [29] describe advanced oxidation processes (AOPs) as aqueous phase oxidation methods. These AOPs are characterized by the use of high energy and the generation of hydroxyl radical (°OH) as a very strong oxidant [7,89,90,91]. Researchers have repeatedly recognized that AOPs effectively degrade pharmaceuticals in wastewater and other aquatic systems. Photochemical, ozone-based (O3/H2O2, O3/UV), and photo-Fenton processes are generally more effective than ozonation alone due to the photon-initiated carbon–halogen bond cleavage and increased generation of hydroxyl free radicals [92]. In addition, UV or visible irradiation can generate reactive excited states of pharmaceuticals. The main problem associated with AOPs is the generation of toxic oxidation products that can remain in the sanitized water [93]. The literature outlined various types of heterogeneous photocatalysts, mechanisms, synthesis methods of the photocatalyst, the nature of the pharmaceutical compound, and photocatalytic reactor designs and their operating parameters, as well as the effect of the water matrices on the degradation [94,95].
The removal of paracetamol and acetaminophen ACE, as NSAID products, was mentioned by many studies [96,97,98]. Borràs-Ferris et al. [99] and Gomez-Avilés et al. [96] used two totally different photocatalytic processes for the elimination of acetaminophen ACE. In fact, the photo-electrocatalytic (PEC) degradation of ACE using TiO2 nanotubes as photocatalysts demonstrated that the photodegradation of ACE was quicker at pH = 3, and a higher degree of mineralization was achieved [99]. In comparison, analyzing by photocatalysis using mixed Ti–Zr metal–organic frameworks under solar irradiation showed total removal of ACE, and high reusability of the photocatalyst had been noted that facilitated the utilization of this photocatalyst in real wastewater treatment [96]. Among analgesics, paracetamol was mentioned in Table 3 for a full degradation at pH 9 [98]. However, using photo-degradation for paracetamol resulted in 98% of removal under solar irradiation, using ferrioxalate for 2 mg/L as a concentration for 5 min of treatment once in an aqueous solution [97]. According to Jallouli et al. [100]., using TiO2 heterogeneous photocatalysis with UV-LEDs was efficient to remove IBU from pharmaceutical industry wastewater and from ultrapure water in natural pH (between 5 to 5.3) and less efficient in treating the wastewater from the municipal WWTPs. In real urban wastewater, they found a percentage between 89.83% and 100% of degradation after 40 min of treatment. Monteoliva-Garcia et al. [101] used light-driven AOPs, which is a good treatment process for three different pharmaceuticals products: ibuprofen, carbamezapine, and ciprofloxacin in real urban wastewater. They demonstrated a removal of 80.4% to 100% of carbamazepine, an antidepressant drug, after 40 min of treatment in real urban wastewater, also from 89.83% to 100% for ibuprofen and a total removal for ciprofloxacin. He et al. [102] also noted the efficiency of the photocatalysis process over photolysis, which was faster and characterized by a higher elimination of pharmaceutical products in wastewater: 74% removal of carbamazepine and 100% removal of both propranolol and DIC. For the other NSAIDs, Calza et al. [103] also studied aqueous solution by TiO2 photocatalyst. They showed a total degradation and mineralization of DIC after 2 h of irradiation. Moreover, for NPX pharmaceutical compounds, Jallouli et al. [104] proved that their removal in ultrapure water by a photocatalytic process is more efficient than photolysis.
In fact, the TiO2-UV process results in higher removals of NPX with 98% and chemical oxygen demand COD of 25%. However, for photolysis at pH initial 6.5, the COD was 83% and 11% after 3 h, whereas Méndez-Arriaga et al. [23] showed the opposite in aqueous solution of 0.8 mmol/L; from their analysis, they noted that 3 h of photolysis process treatment resulted in 90% removal of NPX with just 5% of mineralization, but under the TiO2 photocatalytic process, only 40% of removal and more than 20% mineralization resulted. As for antibiotics, many studies were observed in Table 3, [97,101,105,106,107,108,109,110,111,112]; they all agree on the efficiency of the photocatalytic process for the removal of different antibiotics drugs such as Ciptofloxacin CIP, which was completely removed from urban wastewater even before 20 min of treatment when they added Fe2+ and TiO2 as catalysts in the wastewater treatment plants [101]. Additionally, this drug was eliminated under UVA light by combining two catalysts: TiO2 and ZnO [109]. In their study, they have shown the efficiency of this combination by removal of 300 µg L−1 of CIP from ultra-pure water in less than 6 min. In addition to that, Elmolla and Chaudhuri [107] noted that the removal of antibiotics drugs, amoxicillin, ampicillin, and cloxacillin, under UVA irradiation and zinc oxide was observed when the pH of solution achieved 11. For amoxicillin AMX, it was also analyzed by two other methods, which are identified by Chen et al. [106]. They observed that AMX was completely removed after 90 min of treatment using carbon quantum dots modified potassium titanate nanotubes (CQDs/K2Ti6O13) under UV irradiation (between 365 and 385 nm), and Tran et al. [110] reported a percentage of elimination of AMX and metronidazole of more than 70% after 120 min of treatment by UV/TiO2 photocatalysis at pH medium.
Molinari et al. [113] showed the importance of the photocatalyst process, mainly in the removal of furosemide and ranitidine drugs. In fact, in darkness and without photoreaction, the percentages of rejection were between 10% and 60% for furosemide and between 5% and 30% for ranitidine, but a decrease to 0 was observed in the presence of light, photocatalyst, and oxygen. Despite the high cost of this process, many researchers are adapting these methods because of their effectiveness in the removal of most pharmaceutical drugs: antibiotics, analgesics, NSAIDs, anticancer, beta-blockers, antiseptics, etc. (Table 3). Calza et al. [103] identified the efficiency of the removal of two anticancer drugs, methotrexate and doxorubicin, in ultrapure water, reaching 100% using the UVA/TiO2 system. De la Cruz et al. [114] reported the removal of propranolol in ultrapure water in 240 up to 81% at the solar plant and 94% at the laboratory. From these different studies, we can observe the performance of photocatalysis in degrading pharmaceuticals that is significant. This technique is also characterized by utilizing TiO2, which is inexpensive, with low toxicity and high stability to light illumination, as well as its capacity to be regenerated several times without significantly losing its activity [115].
The antibiotic cloxacillin was also almost completely removed from a synthetic pharmaceutical wastewater using the combined TiO2/photo-Fenton reaction [116]. In addition, the use of a UV-A-LED-photo-Fenton reaction for the removal of antipyrine by Davididou et al. [117] showed a total degradation after 2.5 min and a 93% of mineralization after 60 min. Bautitz and Nogueira [105] have shown the efficiency of the photo-Fenton process for the degradation of tetracycline (TC) in WWTP effluent and surface water (Table 3). They have shown that total removal of TC drug was achieved after 1 min under black light radiation in the presence of Fe(NO3)3 for a concentration of 5.5 mg/L, while solar light irradiation favors ferrioxalate for a 2 mg/L of concentration.
In addition to the different methods of AOPs, ozone treatment enhances the removal of all drugs, including carbamazepine, where removals up to 96% are observed. Ozone is a powerful oxidizing agent (E0 = 2.07 V), capable of acting both directly and indirectly. The direct use of ozone without catalyst or irradiation is ozonation, while the additional use of a catalyst or photoactivation is considered AOP [1]. The removal of drugs such as antibiotics, estrogens, and neutral pharmaceuticals has been treated with Almomani et al. [118] in different types of water for the same pH = 8 and a temperature of 20 °C. For a synthetic wastewater, they found more than 90% of removal for all pharmaceutical products in 2 min with concentration of ozone of 5.5 mg/L. Regarding wastewater, surface water, and wastewater treatment plants, WWTPs, the same percentage of more than 90% removal for antibiotics, estrogens, and neutral pharmaceuticals was achieved with ozone doses of 2.05, 1.11, and 1.30 mg O3/mg DOC, respectively. Moreover, the simple use of ozone for water treatment shows good results, which were confirmed by Dantas et al. [119]. They noted that a total removal of Bezafibrate BZF drugs in the aqueous solution with the initial concentration of 0.5 mmol/L could be reached just after 10 min of treatment using the dose of 0.73 mmol/L of ozone (Table 3). However, degradation almost never results in complete mineralization. In addition, the oxidation of endocrine-distrupting compounds EDCs with comparable chemical structures such as Bisphenol A, 17β-Estradiol, and 17α-Ethynyl Estradiol have been studied by Alum et al. [120]. They reported that, when using ozone at the concentration of 30 mmol/L, 99% transformation would be expected in less than 2 s. However, a residual estrogenic response could still be present because of the oxidation byproducts. They showed a transformation after 10 min. Carballa et al. [121] were interested in the effect of an oxidative pre-treatment with ozone, dose of 20 mg O3 /g TSS, on the removal of pharmaceutical and personal care products (PPCPs) during the anaerobic digestion of sewage sludge. They found an almost total elimination of all PPCPs of 99%, except for carbamazepine, which was only up to 60% removed under thermophilic conditions. Therefore, they concluded that ozonation could represent an economically feasible solution to improve sludge stabilization. Therefore, non-steroidal anti-inflammatory drugs such as indomethacin were removed by Zhao et al. [122] after just 7 min with the lowest dose of ozone.
Concerning the antineoplastic drug 5-fluorouracil, Koltsakidou et al. [123] also studied its degradation under solar irradiation. In fact, the results showed that the ferrioxalate system was more efficient for this degradation and mineralization.

5. Various Methods for the Removal of Pharmaceuticals from Waters

The removal of pharmaceuticals depends on different water treatment processes that result in different efficiencies. Wastewater treatment plants are not designed to completely remove pharmaceuticals [1]. Adsorption has advantages over other methods because of its simple design that can involve low investment in terms of both initial cost and space required [124]. Adsorption processing for water remediation includes low capital investment, applicability at low adsorbate concentrations, suitability for batch or continuous processing, and the ability to reuse and regenerate adsorbents. Adsorption is a low consuming energy process that can be very efficient and lead to a removal of up to 90%, but it has a mild operation condition. With the Fenton process, it is possible to treat large volumes of wastewaters (need of large electrodes or cell stacks). The greatest advantage of photocatalysis lies in the use of the semiconductor TiO2, which is non-hazardous, eco-friendly, inexpensive, and stable [94]. Indeed, in a quick and very high percentage of organic matter, degradation can be reached and can be more remarkable under sunlight irradiation and using in situ cathodic production of H2O2. However, this process presents many disadvantages such as the need for acidic conditions with a pH value near to 3.0. Thus, a neutralization is requested after the treatment, leading to the production of a large amount of sludge. Special attention should be paid to the formation of halogenated byproducts. According to Patel et al. [1], an advanced oxidation process can lead to a high degradation of pharmaceuticals, but it is difficult to apply on a wide scale of WWTPs due to the oxidative byproducts and high operation cost. A common problem to all AOPs is their high cost, mainly because of the high demand for electrical energy (Figure 2).
Table 3. Removal of different pharmaceutical products by AOPs.
Table 3. Removal of different pharmaceutical products by AOPs.
Operation ProcessDrugsAOPsScaleMatrixExperiment Duration (min)Reactor Volume (mL)Significant FindingsReferences
PhotocatalysisAcetaminophen ACEMixed Ti-Zr metal-organic-frameworksLab-scaleAqueous solution18015090% of removal for ACE.[96]
TiO2 nanotubes photo catalystsLab-scaleAqueous solution12012The maximum conversion value for elimination was reached at pH = 3[99]
ParacetamolTiO2 nanoparticles and TiO2/ under UV and sunlight irradiationLab-scaleUltrapure water1501500More than 90% of degradation
pH optimum = 9
[98]
AspirinTiO2-based nanosheetsLab-scaleAqueous solution21030100% of degradation
87.8% after 120 min
[125]
Ibuprofen IBUTiO2/ UV-LEDLab-scaleUltrapure water, municipal wastewater,
pharmaceutical industry
wastewater
3025042% of removal for ultrapure water, 18% degradation for municipal wastewater and only 9% removal for PIWW[100]
DiclofenacTiO2 suspensionsLab-scaleAqueous solution12050100% of degradation[103]
Azithromycin
Trimethoprim
Ofloxacin
Sulfamethoxazole
TiO2/ UVA-LEDsLab-scaleSecondary urban wastewater60150Total degradation[126]
Metronidiazole
Amoxicillin
UV/TiO2Lab-scaleAqueous solution1201000More than 70% of degradation[110]
MetronidiazoleUV/TiO2Lab-scaleWater matrix301000Degradation of MNZ decreased in the presence of H2PO4, Fe3+, and humic acid (HA)[111]
CiptoflaxacinTiO2/UVA
ZnO/ UVA
Lab-scaleUltra-pure water650Efficiency of degradation of CIP for two catalysts[109]
AmoxicillinCarbon quantum dots modified K2Ti6O13 nanotubes/ UVALab-scaleDeionized water9050Completely removal of AMX.[106]
Amoxicillin
Ampicillin
Cloxacillin
UV/ZnOLab-scaleAqueous solution180500Total degradation for pH = 11.[107]
Penicillin GTi3+ self-doped TiO2Lab-scaleMQ-water1205098.3% elimination of penicillin G.[108]
Furosemide Ranitidine Ofloxacine Phenazone
Naproxen
Carbamazepine Clofibric acid
TiO2Lab-scaleAqueous solution900200Degradation of all pharmaceutical products.[113]
CarboplatinTiO2/UVLab-scaleAqueous solution30250100% of degradation[127]
Ifosfamide CyclophosphamidePt-doped TiO2
TiO2
Lab-scaleAqueous solution2405092% removal for cyclophosphamide and 95% removal for ifosfamide.[128]
CytarabineTiO2/ solar lightLab-scaleAqueous solution360100Cytarabine was degraded in all experimental conditions.[123]
Methotrexate DoxorubicinTiO2 Degussa P25Lab-scaleUltrapure water60
15
560% of removal for methotrexate
43% of removal for doxorubicin
[129]
Chlorhexidine digluconateTiO2Technical scaleAqueous solution6010099% of removal.[130]
Carbamazepine
Gabapentin
Lamotrigine
Oxcarbazepine
Venlafaxine
Bisoprolol
Celiprolol
Talinolol
Bezafibrate
Tramadol
Candesartan
Eprosartan
Irbesartan
Valsartan
ZnO and TiO2 under UVA radiationLab-scaleWastewater treatment plants30100ZnO shows a higher degradation of pharmaceutical products than TiO2[131]
Photolysis and photocatalysisPropranolol
Carbamazepine
Diclofenac
TiO2/solar lightLab-scaleWastewater effluent5760500100% for propranolol, 100% for diclofenac;
76 ± 3% for carbamazepine
[102]
NaproxenDirect photolysis and TiO2/UVLab-scaleAqueous solution18012083% after photolysis
98% after photo catalysis
[104]
Direct photolysis/ TiO2/UVLab-scaleAqueous solution1803090% after photolysis
40% after photo catalysis
[23]
PropranololDirect photolysisTechnical- scaleUltrapure water2401000071% of removal[114]
Lab-scale77% of removal
TiO2Technical-scale 100081% of removal
Lab-scale94% of removal
Oxolinic acidTiO2Lab-scaleUltrapure water30100TiO2 photocatalysis can be an efficient methods and rapid way for the elimination of oxolinic acid.[132]
Photo-Fenton and photocatalysisIbuprofen
Carbamazepine Ciprofloxacin
Photo-assisted
H2O2/ UV AOP
Lab-scaleReal urban wastewater40
20
70089.83 and 100% degradation for IBU
80.14 to 100% for carbamazepine
100% removal for CIP
[101]
AntipyrineUV-A-LED-photo Fenton reactionLab-scaleAqueous solution2.5150100% of degradation after 2.5 min.
93% of mineralization at 60 min.
[117]
CloxacillinTiO2/ photo-Fenton processLab-scaleSynthetic pharmaceutical wastewater240150Total degradation for antibiotic.[116]
Photo-FentonTetracyclineFe(NO3)3/ solar and black light irradiationLab-scaleSTP effluents1.5500Total degradation of TC.[105]
Amoxicillin
Bezafibrate
Paracetamol
H2O2/ solar and black light irradiationLab-scaleAqueous solution1800With FeOx 84% removal for amoxicillin or 62% was observed using Fe(NO3)3;
98% of removal for bezafibrate and paracetamol in the presence of Fe(NO3)3.
[97]
5-fluorouracil(Fe3+/H2O2)/
([Fe(C2O4)3] 3− /H2O2, Fe3+/S2O8 2−)
Lab-scaleUltrapure water20–60100In this study, it was found that the ferrioxalate system was more efficient for the degradation and mineralization of 5- fluorouracil.[133]
OzonationAntibiotics, steroid hormone, lipid regulator, antineoplastic, non-steroidal anti-inflammatory drug, and psychostimulantDose of ozoneLab-scaleSynthetic wastewater, surface water, and
the effluent of wastewater treatment plant
12000The optimum dose of ozone was found to be yielding of 118.1, 222.3, and 222.4 mg/h, respectively, for synthetic waste water, surface water, and effluent of wastewater treatment plants.
Ozonation experiments were carried out at 20 °C and at a pH of 8.
This experience eliminated >99.9% of removal for most of the studied pharmaceuticals.
The increased toxicity for aqueous solutions of acidic pharmaceuticals at a specific ozone dose of 2.24 mg O3/mg DOC was due to formation of more toxic byproducts.
[118]
IndomethacinOzone and hydroxyl radicalsLab-scaleUltrapure water7300This drug was eliminated within 7 min under the lowest ozone dose, but TOC removal was only 50% even under the highest ozone dose used in the experiments[122]
Amoxicillinmedium-high frequency ultrasonic irradiation and/or ozonationLab-scaleAqueous solution90250The highest removal was achieved at 575 kHz ultrasonic frequency (>99%) with the highest pseudo first order reaction rate constant 0.04 min at pH 10.[134]
BezafibrateDose of ozoneLab-scaleAqueous solution10800Total degradation after 10 min with a dose of ozone = 0.73 mmol/L[119]
Bisphenol A
17β-Estradiol,
17α-Ethynyl Estradiol
OzoneLab-scaleAqueous solution1040Total removal for drugs and a rapid transformation in 10 min.[120]
Pharmaceutical and personal care products (PPCPs)Ozonelab-scaleAqueous solution12068099% of elimination for all pharmaceuticals products.[121]

6. AOPs/Adsorption as a Promising Alternative Treatment Technique

Many studies have shown the negative effects of PPCPs on the environment that we have mentioned before (3), essentially the endocrine-disrupting effects and the risk of the spread of antibiotic resistance [135]. According to the OECD [136], many countries have imposed rules to prevent the risk of developing drugs in the environment. For example, the United States has national regulations on the disposal of hazardous pharmaceutical waste in the health sector, and Germany has developed an environmental checklist for veterinarians and farmers with the aim of reducing the use and release of veterinary pharmaceuticals to the environment. Additionally, Korea uses suspect and non-target screening to identify and prioritize pharmaceuticals for water quality monitoring. In the same order, Australia has a national pharmaceutical collection and disposal program.
In this review, we show clearly that one singular method such as conventional treatment, physical treatment methods such as coagulation and adsorption, or advanced oxidation ones, including a combination of oxidizing agents (such as H2O2 and O3), irradiation (such as ultraviolet light or ultrasound), and catalysts (such as Fe2+), cannot remove them sufficiently. Thus, a combination of two or more processes is required to secure the total elimination of drugs in water.
The simultaneous application of ozonation and adsorption for the removal of imidazole, reported by Patel et al. [1], shows that 90–100% of degradation was achieved by ozonation and 10–20% of mineralization. The presence of an adsorbent further reduces toxicity by 30%. Catalytic ozonation associated with adsorption can be effective for the removal of pharmaceuticals. In fact, ozonation degraded the drug products and the adsorbent can remove any remaining pharmaceutical and degradation byproducts. Additionally, Mojiri et al. [137] show that the application of a combined process of ozone and adsorption (cross-linked chitosan/bentonite) for the treatment of solutions containing acetaminophen and amoxicillin was efficiently removed by 84.8% and 82.7%, with an initial concentration of ozone of 0.17 and 0.16 mg/L, respectively (Table 4). Combined photo-Fenton and biological treatment were over 95% efficient, of which 33% corresponds to the solar photochemical process and 62% to the biological treatment [138]. Additionally, Sulfamethoxazole was successfully removed using both catalytic ozonation and adsorption on modified powdered activated carbon (Fe2O3/CeO2-loaded activated carbon) at 2 g/L, applied for only 50 min [139]. Sui et al. [140] studied the removal of 13 pharmaceutical and personal care products (PPCPs) by sequential ultraviolet and ozonation process in a full-scale wastewater treatment plant (Table 4). Authors showed that most of the target PPCPs were effectively removed, and the median removal efficiencies of individual PPCPs, ranging from 13% to 89%, were dependent on their reaction rate constants with molecular ozone. Moreover, Salgado et al. [141] studied the removal of eighteen drugs that have a biological removal rate in a full-scale activated sludge plant combined with UV photolysis. They showed that up to 75% of removal was usually found for 17 of the 18 most commonly detected PPCPs, with the only exception being for diclofenac, which has often shown negative values for the rate of biological elimination in activated sludge and has been mainly degraded by UV photolysis. In fact, in treatment water, there are different possibilities of combining advanced oxidation processes such as photo-Fenton/ozonation and photocatalysis/ozonation [142,143]. At the end of the treatment, the combined processes could generate transformation products more persistent and more toxic than the original pharmaceutical products. The combined processes AOPs/adsorption provided almost complete removal of pollution. Therefore, adsorption can remove low-molecular organic products issued from AOPs to achieve more cost-effective and advanced treatment. Thus, integration of AOP and adsorption has been proposed by many researchers to solve the shortcomings experienced when these two technologies are applied separately [144]. From this review, we can highlight that among the combined processes reported, the combination of photocatalysis and adsorption seems more attractive and has been presented as the most effective solution for the removal of pharmaceutical compounds and their byproducts from water. Many studies have shown the efficiency of combination process, such as [145,146,147]. In fact, Brienza et al. [145] used methods of combination micelle-clay sorption to solar photo-Fenton in different pharmaceutical products in almost 13 drugs: Sulfamethazine, Caffeine, Tamoxifen, Ketoprofen, Sulfamethoxazole, Mepanypirim Diclofenac, Chlotianidin, Amoxicillin, Venlafaxine, Fenofibric acid, Carbamazepine, and Atenolol. They took the GAC from Aquacarb (Chemivor Carbon Company, city country), sand of Fontainebleau from PROLABO (particle size 150–210 μm, France), and montmorillonite from Wyoming Namontmorillonite SWγ-2 (grain size 0.8–1.5 mm, Montmorillonite Minerals Society, Columbia, MO). The lowest percentage of removal was tamosifen with 37.71%, contrary to others that were almost completely degraded. In Table 4, Rahmani et al. [147] used a combination of electrooxidation–ozonation for the antibiotic ciprofloxacin. This combination shows high performance for elimination about 90% in just 150 min.
Looking for the performance of the combination processes for adsorption/photocatalysis, both Lin et al. [148] and Wang et al. [146] used TiO2 nanofibers combined with BN nanosheets and BiOClxIy nanospheres, respectively. Lin et al. [148] found a higher degradation for ibuprofen. Additionally, Wang et al. [146] studied two different drugs, acetaminophen and hydroxyphenylacteric-acid (p-HPA). They resulted in a total removal for p-HPA and an 80% elimination for acetaminophen.
In fact, when pharmaceutical products become very harmful and our groundwater is at risk of contamination, we must become involved, like Lhotsky et al. [149]. They noted the presence of pharmaceutical plants for more than 80 years in the Czech Republic. For these reasons, they took the three principal dangerous drugs, benzene, toluene, and chlorobenzene, found in the site. They resulted from a higher elimination of these products after the treatment combination of UV/H2O2 for a 99% of removal.
Table 4. Removal of different drugs by combined processes.
Table 4. Removal of different drugs by combined processes.
Operation ProcessDrugsCombining ProcessScaleMatrixExperience Time (Min)Reactor Volume (mL)Significant FindingsReferences
Ozone oxidation combined to adsorptionAcetaminophen
Amoxicillin
Ozone gas diffusionTechnical scaleAqueous solution25210084.8% of removal
82.7% of removal
[137]
Acetaminophen
Amoxicillin
Cross-linked
chitosan/bentonite
25200Total removal of two drugs
SulfamethoxazolePAC and Fe2O3/CeO2 loaded activated carbon MOPACLab-scaleAqueous solution30100Sulfamethoxazole was highly removed by combining adsorption activated carbon MOPAC with ozonation at pH = 3.5.[139]
Ozone5250
Photocatalysis Combined to adsorptionCiptoflaxacinGraphitized mesoporous carbonLab-scaleAqueous solution10100Totally mineralized in 1.5 h.[112]
TiO220200
UV/O3 process combined to adsorptionBizafibrate, clofibric acid, carbamezapine, caffeine, chloramphenicol, N,Ndiethyl-meta-tolumide, diclofenac, gemfibrozil, mefenamic acid, metoprolol, propanolol, sulpiride, trimethoprimUV and ozonationTechnical-scaleWWTPs406DF, TP, CP, and CBZ were removed about 80%.
SP, MA, CA, PNN, and MTP are efficiently removed.
[140]
Photolysis combined to adsorptionDiclofenac, Etafenamete, Ibuprofen, ketoprofen, fluoxetine, clorazepate, hydroxyzine, indapamide, analapril, captopril, atenolol, cloridric-acid, ampicillinSludge adsorptionTechnical-scaleWastewater treatment plant2 weeks600 m3Incomplete removal of diclofenac.
Removal of pharmaceuticals by adsorption was about 33% and UV was 22%.
[141]
UV radiation with hydrogen peroxidebenzene, toluene and chlorobenzeneUV/H2O2 with aerationTechnical-scaleUltrapure water4 months1000High efficiency of both techniques for the removal of drugs.
Removal efficiency ringing from 72% to 99%.
[149]
Electrooxidation-ozonationCiprofloxacinTi/PbO2Technical-scaleAqueous solution9028090% of removal[147]
Photo-Fenton combined to adsorptionSulfamethazine
Caffeine,
Tamoxifen
Ketoprofen
Sulfamethoxazole
Mepanypirim
Diclofenac
Chlotianidin
Amoxicillin
Venlafaxine
Fenofibric acid
Carbamazepine
Atenolol
GAC/photo-FentonLab-scale
Technical-scale
Biological domestic wastewater treatment plant120600087.47% of removal
93.64% of removal
37.91% of removal
100% of removal
99.98% of removal
100% of removal
96.67% of removal
100% of removal
100% of removal
100% of removal
100% of removal
99.97% of removal
100% of removal
[145]
Photocatalysis combined to adsorptionIbuprofenTiO2/boron nitride (BN) nanocompositesLab-scaleAqueous solution90100Efficiency of the combination for the removal of ibuprofen[148]
Acetaminophen and hydroxyphenylacetic acid (p-HPA)BiOClxIy nanospheresLab-scaleAqueous solution18050100% removal of p-HPA and 80% removal of acetaminophen.[146]
Persulfate oxidation process combined to adsorptionsulfadimethoxine (SDM), sulfamonomethoxine (SMM), and sulfachloropyridazine (SCP)MIL-101(Cr)Lab-scaleDeepwater901000The adsorption capacity of MIL-101(Cr) decreased with the increase of oxidation times.
The maximum adsorption capacities to SDM, SMM, and SCP were 588, 196, and 196 mg/g, respectively, by Langmuir at 25 °C
[150]
Ozonation and sonolysisDiclofenac, Sulfamethoxazole, CarbamazepineOzonation and ultrasound (O3/US)Lab-scaleWastewater1208000The remove efficiency yielding from 80% to 84%[151]

7. Conclusions, Future Perspectives, and Challenges

The large-scale use of hospital, household, and industrial wastewaters has led to widespread contamination of water by different hazardous drugs, increasing concern for human health and the environment. Various analyses have been tried to reduce the presence of drugs in water to save the environment and the human body. For the removal of pharmaceutical products from water, many methods were taken into consideration, such as adsorption and advanced oxidation processes. In this review, we demonstrated that the application of one technology to treat pharmaceutical wastewaters seems insufficient for the total removal of PPCPs from WWTPs before discharge. In fact, a trace of these PPCPs has also been found in rivers and lakes. An advanced oxidation process has been developed and constructed to show promising potential as a future generation for the treatment of water. The combination of adsorption and AOP processes can be considered an excellent way to completely remove pharmaceutical products such as antibiotics, NSAIDs, and analgesics, even at trace concentrations. A variety of techniques have been taken into account in this review to reduce the toxicity of PPCPs in water on humans via fresh drinkable water or food as well as on the environment (lakes and rivers).
This review also highlights the combination of the different treatment methods to achieve a high percentage of PPCPs removal that reaches up to 100% in a short operation time. In the case of coupling AOPs to adsorption, a number of low-cost and local materials such as clay, activated carbon, and olive waste can be used as adsorbents. Moreover, the operation costs and sustainability of this kind of process have to be optimized, notably by using renewable energy, i.e., solar light instead of artificial illumination.

Author Contributions

Conceptualization, M.K. and N.R.; methodology, M.K.; writing—original draft preparation F.M.; writing—reviewing K.C.; supervision, M.K. and N.R.; project administration, M.K.; funding acquisition, N.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All authors ensure that data shared are in accordance with consent provided by participants on the use of confidential data.

Acknowledgments

The authors gratefully acknowledge the support of the University of Sfax in Tunisia and Aix-Marseille University in France.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

AC:activated carbon
ACE:acetaminophen
AMX:amoxicillin
AOPs:advanced oxidation process
AUJ:Aurajok
AZT:azithromycin
B:bentonite
BF:Bezafibrate
CA:Clofibric acid
CBZ:carbamezapine
CF:Caffeine
CIP:Ciptoflaxacin
COD:Chemical oxygen demand
CP:Chloramphenicol
CTAC:Cyclamen persicum tubers based activated carbon
CTC:Chloitetracycline-HCI
CuNPs:Copper nanoparticles
DEET:N,N-diethyl-meta-tolumide
DIC:Diclofenac
DMC:Democlocycline-HCI
DXC:Doxycycline-hyclate
EDCs:Endocrine disruption coumpounds
GAC:Granular activated carbon
Gf:Gemfibrozil
IBU:Ibuprofen
K:kaolinite
Kyj:kyronjoki
KOJ:Kokemäenjoki
MA:Mefenamic acid
MCC:Meclocycline-sulfosalicylate
MDD:major depressive disorder
MNC:Minocycline-HCI
MNZ:metronidiazole
MT:Montmorillonite
MTP:Metoprolol
MPs:Microplastics
NPX:Naproxen
NSAIDs:Non-Steroidalanti-inflammatory drugs
OCD:obsessive-compulsive disorder
OFX:oflaxacin
OTC:Oxytetracycline-HCI
PACI:Poly-aluminuim chloride
PEC:Photo-electrocatalytic
PMDD:premenstrual dysphoric disorder
PPCPs:Pharmaceuticals and personal care products
PPN:Propanolol
SEJ:Seinajok
SMX:Sulfamethoxazole
SP:Sulpiride
TAs:tetracycline classes of antibiotics
TC:tetracycline
TMP:trimethoprim
UK:united Kingdom
US:united States
UVA:ultraviolet A
VER:vermiculite
WW:waste water
WWT:waste water treatment
WWTP:waste water treatment plants

References

  1. Patel, M.; Kumar, R.; Kishor, K.; Mlsna, T.; Pittman, C.U.; Mohan, D. Pharmaceuticals of Emerging Concern in Aquatic Systems: Chemistry, Occurrence, Effects, and Removal Methods. Chem. Rev. 2019, 119, 3510–3673. [Google Scholar] [CrossRef] [Green Version]
  2. Jodeh, S.; Abdelwahab, F.; Jaradat, N.; Warad, I.; Jodeh, W. Adsorption of diclofenac from aqueous solution using Cyclamen persicum tubers based activated carbon (CTAC). J. Assoc. Arab. Univ. Basic Appl. Sci. 2016, 20, 32–38. [Google Scholar] [CrossRef] [Green Version]
  3. Kalyva, M. Fate of Pharmaceuticals in the Environment—A Review; Dept. of Ecology and Environmental Science (EMG) S-901 87 Umeå: Vasterbotten County, Sweden, 2017; p. 30. [Google Scholar]
  4. Roberts, P.; Thomas, K. The occurrence of selected pharmaceuticals in wastewater effluent and surface waters of the lower Tyne catchment. Sci. Total. Environ. 2006, 356, 143–153. [Google Scholar] [CrossRef]
  5. Wang, J.; Chu, L.; Wojnárovits, L.; Takács, E. Occurrence and fate of antibiotics, antibiotic resistant genes (ARGs) and antibiotic resistant bacteria (ARB) in municipal wastewater treatment plant: An overview. Sci. Total. Environ. 2020, 744, 140997. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, J.; Wang, S. Removal of pharmaceuticals and personal care products (PPCPs) from wastewater: A review. J. Environ. Manag. 2016, 182, 620–640. [Google Scholar] [CrossRef]
  7. Szabó, R.K.; Megyeri, C.; Illés, E.; Gajda-Schrantz, K.; Mazellier, P.; Dombi, A. Phototransformation of ibuprofen and ketoprofen in aqueous solutions. Chemosphere 2011, 84, 1658–1663. [Google Scholar] [CrossRef] [PubMed]
  8. Mansouri, H.; Carmona, R.J.; Gomis-Berenguer, A.; Souissi-Najar, S.; Ouederni, A.; Ania, C.O. Competitive adsorption of ibuprofen and amoxicillin mixtures from aqueous solution on activated carbons. J. Colloid Interface Sci. 2015, 449, 252–260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Petrovic, M.; de Alda, M.J.L.; Diaz-Cruz, S.; Postigo, C.; Radjenovic, J.; Gros, M.; Barcelo, D. Fate and removal of pharmaceuticals and illicit drugs in conventional and membrane bioreactor wastewater treatment plants and by riverbank filtration. Philos. Trans. R. Soc. Math. Phys. Eng. Sci. 2009, 367, 3979–4003. [Google Scholar] [CrossRef]
  10. Castiglioni, S.; Bagnati, R.; Fanelli, R.; Pomati, F.; Calamari, D.; Zuccato, E. Removal of Pharmaceuticals in Sewage Treatment Plants in Italy. Environ. Sci. Technol. 2006, 40, 357–363. [Google Scholar] [CrossRef]
  11. Klein, E.Y.; van Boeckel, T.P.; Martinez, E.M.; Pant, S.; Gandra, S.; Levin, S.A.; Goossens, H.; Laxminarayan, R. Global increase and geographic convergence in antibiotic consumption between 2000 and 2015. Proc. Natl. Acad. Sci. USA 2018, 115, E3463–E3470. [Google Scholar] [CrossRef] [Green Version]
  12. Wang, J.; Zhuan, R.; Chu, L. The occurrence, distribution and degradation of antibiotics by ionizing radiation: An overview. Sci. Total. Environ. 2019, 646, 1385–1397. [Google Scholar] [CrossRef]
  13. Fatta-Kassinos, D.; Hapeshi, E.; Achilleos, A.; Meric, S.; Gros, M.; Petrovic, M.; Barcelo, D. Existence of Pharmaceutical Compounds in Tertiary Treated Urban Wastewater that is Utilized for Reuse Applications. Water Resour. Manag. 2011, 25, 1183–1193. [Google Scholar] [CrossRef]
  14. Gracia-Lor, E.; Sancho, J.V.; Serrano, R.; Hernández, F. Occurrence and removal of pharmaceuticals in wastewater treatment plants at the Spanish Mediterranean area of Valencia. Chemosphere 2012, 87, 453–462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Mirzaei, R.; Yuesian, M.; Nasseri, S.; Gholami, M.; Jalilzadeh, E.; Shoeibi, S.; Mesdaghinia, A. Occurrence and fate of most prescribed antibiotics in different water environments of Tehran, Iran. Sci. Total Environ. 2018, 619–620, 446–459. [Google Scholar] [CrossRef]
  16. Xiang, J.; Wu, M.; Lei, J.; Fu, C.; Gu, J.; Xu, G. The fate and risk assessment of psychiatric pharmaceuticals from psychiatric hospital effluent. Ecotoxicol. Environ. Saf. 2018, 150, 289–296. [Google Scholar] [CrossRef]
  17. Silva, A.L.P.; Prata, L.C.; Walker, T.R.; Duarte, A.C.; Ouyang, W.; Barcelò, D.; Rocha-Santos, T. Increased plastic pollution due to Covid-19 pandemic: Challenges and recommendations. Chem. Eng. J. 2020, 405, 126683. [Google Scholar] [CrossRef] [PubMed]
  18. Kalantary, R.R.; Jamshidi, A.; Mofrad, M.M.G.; Jafari, A.J.; Heidari, N.; Fallahizadeh, S.; Arani, M.H.; Torkashvand, J. Effect of COVID-19 pandemic on medical waste management: A case study. J. Environ. Health Sci. Eng. 2021, 19, 831–836. [Google Scholar] [CrossRef]
  19. Haque, M.d.S.; Uddin, S.; Sayem, S.M.; Mohib, K.M. Coronavirus disease 2019 (COVID-19) induced waste scenario: A short overview. J. Environ. Chem. Eng. 2021, 9, 104660. [Google Scholar] [CrossRef]
  20. El Majid, B.; El Hammoumi, A.; Motahhir, S.; Lebbadi, A.; El Ghzizal, A. Preliminary design of an innovative, simple, and easy-to-build portable ventilator for COVID-19 patients. Euro-Mediterr. J. Environ. Integr. 2020, 5, 23. [Google Scholar] [CrossRef]
  21. Atugoda, T.; Vithanage, M.; Wijesekara, H.; Bolan, N.; Sarmah, A.K.; Bank, M.S.; You, S.; Ok, Y.S. Interactions between microplastics, pharmaceuticals and personal care products: Implications for vector transport. Environ. Int. 2021, 149, 106367. [Google Scholar] [CrossRef]
  22. Fu, Y.; Gao, X.; Geng, J.; Li, S.; Wu, G.; Ren, H. Degradation of three nonsteroidal anti-inflammatory drugs by UV/persulfate: Degradation mechanisms, efficiency in effluents disposal. Chem. Eng. J. 2019, 356, 1032–1041. [Google Scholar] [CrossRef]
  23. Méndez-Arriaga, F.; Gimenez, J.; Esplugas, S. Photolysis and TiO2 Photocatalytic Treatment of Naproxen: Degradation, Mineralization, Intermediates and Toxicity. J. Adv. Oxid. Technol. 2008, 11, 435–444. [Google Scholar] [CrossRef]
  24. Manrique-Moreno, M.; Suwalsky, M.; Villena, F.; Garidel, P. Effects of the nonsteroidal anti-inflammatory drug naproxen on human erythrocytes and on cell membrane molecular models. Biophys. Chem. 2010, 147, 53–58. [Google Scholar] [CrossRef]
  25. Ghauch, A.; Tuqan, A.M.; Kibbi, N. Naproxen abatement by thermally activated persulfate in aqueous systems. Chem. Eng. J. 2015, 279, 861–873. [Google Scholar] [CrossRef]
  26. Lubet, R.A.; Scheiman, J.M.; Bode, A.; White, J.; Minasian, L.; Juliana, M.M.; Grubbs, C.J. Prevention of Chemically Induced Urinary Bladder Cancers by Naproxen: Protocols to Reduce Gastric Toxicity in Humans Do Not Alter Preventive Efficacy. Cancer Prev. Res. (Phila. Pa.) 2015, 8, 296–302. [Google Scholar] [CrossRef] [Green Version]
  27. Moslah, B.; Hapeshi, E.; Jrad, A.; Fatta-Kassinos, D.; Hedhili, A. Pharmaceuticals and illicit drugs in wastewater samples in north-eastern Tunisia. Environ. Sci. Pollut. Res. 2018, 25, 18226–18241. [Google Scholar] [CrossRef] [PubMed]
  28. Park, S.; Rhee, M.-Y.; Lee, S.Y.; Park, S.W.; Jeon, D.; Kim, B.-W.; Kwan, J.; Choi, D. A prospective, randomized, open-label, active-controlled, clinical trial to assess central haemodynamic effects of bisoprolol and atenolol in hypertensive patients. J. Hypertens. 2013, 31, 813–819. [Google Scholar] [CrossRef]
  29. Klavarioti, M.; Mantzavinos, D.; Kassinos, D. Removal of residual pharmaceuticals from aqueous systems by advanced oxidation processes. Environ. Int. 2009, 35, 402–417. [Google Scholar] [CrossRef]
  30. Bártíková, H.; Podlipná, R.; Skálová, L. Veterinary drugs in the environment and their toxicity to plants. Chemosphere 2016, 144, 2290–2301. [Google Scholar] [CrossRef]
  31. Nakada, N.; Tanishima, T.; Shinohara, H.; Kiri, K.; Takada, H. Pharmaceutical chemicals and endocrine disrupters in municipal wastewater in Tokyo and their removal during activated sludge treatment. Water Res. 2006, 40, 3297–3303. [Google Scholar] [CrossRef]
  32. Radjenovic, J.; Petrovic, M.; Barceló, D. Analysis of pharmaceuticals in wastewater and removal using a membrane bioreactor. Anal. Bioanal. Chem. 2007, 387, 1365–1377. [Google Scholar] [CrossRef] [Green Version]
  33. Santos, J.L.; Aparicio, I.; Alonso, E. Occurrence and risk assessment of pharmaceutically active compounds in wastewater treatment plants. A case study: Seville city (Spain). Environ. Int. 2007, 33, 596–601. [Google Scholar] [CrossRef]
  34. Terzić, S.; Senta, I.; Ahel, M.; Gros, M.; Petrovic, M.; Barcelo, D.; Muller, J.; Knepper, T.; Marti, I.; Ventura, F.; et al. Occurrence and fate of emerging wastewater contaminants in Western Balkan Region. Sci. Total Environ. 2008, 399, 66–77. [Google Scholar] [CrossRef] [Green Version]
  35. Akhtar, J.; Amin, N.A.S.; Shahzad, K. A review on removal of pharmaceuticals from water by adsorption. Desalination Water Treat. 2016, 57, 12842–12860. [Google Scholar] [CrossRef]
  36. Göbel, A.; McArdell, C.S.; Suter, M.J.-F.; Giger, W. Trace Determination of Macrolide and Sulfonamide Antimicrobials, a Human Sulfonamide Metabolite, and Trimethoprim in Wastewater Using Liquid Chromatography Coupled to Electrospray Tandem Mass Spectrometry. Anal. Chem. 2004, 76, 4756–4764. [Google Scholar] [CrossRef]
  37. Wang, J.; Wang, S. Microbial degradation of sulfamethoxazole in the environment. Appl. Microbiol. Biotechnol. 2018, 102, 3573–3582. [Google Scholar] [CrossRef]
  38. Puckowski, A.; Cwięk, W.; Mioduszewska, K.; Stepnowski, P.; Białk-Bielińska, A. Sorption of pharmaceuticals on the surface of microplastics. Chemosphere 2021, 263, 127976. [Google Scholar] [CrossRef]
  39. Chouchene, K.; Rocha-Santos, T.; Ksibi, M. Types, occurrence, and distribution of microplastics and metals contamination in sediments from south west of Kerkennah archipelago, Tunisia. Environ. Sci. Pollut. Res. 2020. [Google Scholar] [CrossRef]
  40. Prinz, N.; Korez, Š. Understanding How Microplastics Affect Marine Biota on the Cellular Level Is Important for Assessing Ecosystem Function: A Review. In YOUMARES 9-The Oceans: Our Research, Our Future; Jungblut, S., Liebich, V., Bode-Dalby, M., Eds.; Springer: Berlin/Heidelberg, Germany, 2020; pp. 101–120. [Google Scholar] [CrossRef] [Green Version]
  41. Lin, L.; Tang, S.; Wang, X.S.; Sun, X.; Han, Z.; Chen, Y. Accumulation mechanism of tetracycline hydrochloride from aqueous solutions by nylon microplastics. Environ. Technol. Innov. 2020, 18, 100750. [Google Scholar] [CrossRef]
  42. Christou, A.; Karaolia, P.; Hapeshi, E.; Michael, C.; Fatta-Kassinos, D. Long-term wastewater irrigation of vegetables in real agricultural systems: Concentration of pharmaceuticals in soil, uptake and bioaccumulation in tomato fruits and human health risk assessment. Water Res. 2017, 109, 24–34. [Google Scholar] [CrossRef]
  43. Khan, G.A.; Berglund, B.; Khan, K.M.; Lindgren, P.-E.; Fick, J. Occurrence and Abundance of Antibiotics and Resistance Genes in Rivers, Canal and near Drug Formulation Facilities—A Study in Pakistan. PLoS ONE 2013, 8, e62712. [Google Scholar] [CrossRef]
  44. Kumar, N.N.; Aiyagari, N. Artificial Recharge of Groundwater. Groundwater Pollution Primer. Available online: https://leg.mt.gov/content/Committees/Interim/2005_2006/environmental_quality_council/eetings/minutes/eqc09112006_ex20.pdf (accessed on 7 June 1998).
  45. Skoczko, I.; Puzowski, P.; Szatyłowicz, E. Experience from the Implementation and Operation of the Biological Membrane Reactor (MBR) at the Modernized Wastewater Treatment Plant in Wydminy. Water 2020, 12, 3410. [Google Scholar] [CrossRef]
  46. Liao, B.; Bokhary, A.; Cui, L.; Lin, H. A Review of Membrane Technology for Integrated Forest Biorefinery. J. Membr. Sci. Res. 2017, 3, 120–141. [Google Scholar] [CrossRef]
  47. Petrie, B.; McAdam, E.J.; Scrimshaw, M.D.; Lester, J.N.; Cartmell, E. Fate of drugs during wastewater treatment. TrAC Trends Anal. Chem. 2013, 49, 145–159. [Google Scholar] [CrossRef] [Green Version]
  48. Tauxe-Wuersch, A.; de Alencastro, L.F.; Grandjean, D.; Tarradellas, J. Occurrence of several acidic drugs in sewage treatment plants in Switzerland and risk assessment. Water Res. 2005, 39, 1761–1772. [Google Scholar] [CrossRef]
  49. Lindqvist, N.; Tuhkanen, T.; Kronberg, L. Occurrence of acidic pharmaceuticals in raw and treated sewages and in receiving waters. Water Res. 2005, 39, 2219–2228. [Google Scholar] [CrossRef]
  50. Metcalfe, C.D.; Koenig, B.G.; Bennie, D.T.; Servos, M.; Ternes, T.A.; Hirsch, R. Occurrence Of Neutral And Acidic Drugs In The Effluents Of Canadian Sewage Treatment Plants. Environ. Toxicol. Chem. 2003, 22, 2872. [Google Scholar] [CrossRef]
  51. Clara, M.; Strenn, B.; Gans, O.; Martinez, E.; Kreuzinger, N.; Kroiss, H. Removal of selected pharmaceuticals, fragrances and endocrine disrupting compounds in a membrane bioreactor and conventional wastewater treatment plants. Water Res. 2005, 39, 4797–4807. [Google Scholar] [CrossRef]
  52. Strenn, B.; Clara, M.; Gans, O.; Kreuzinger, N. Carbamazepine, diclofenac, ibuprofen and bezafibrate--investigations on the behaviour of selected pharmaceuticals during wastewater treatment. Water Sci. Technol. 2004, 50, 269–276. [Google Scholar] [CrossRef] [PubMed]
  53. Ternes, T.A.; Meisenheimer, M.; McDowell, D.; Sacher, F.; Brauch, H.-J.; Haist-Gulde, B.; Preuss, G.; Wilme, U.; Zulei-Seibert, N. Removal of Pharmaceuticals during Drinking Water Treatment. Environ. Sci. Technol. 2002, 36, 3855–3863. [Google Scholar] [CrossRef]
  54. Kreuzinger, N.; Clara, M.; Strenn, B.; Kroiss, H. Relevance of the sludge retention time (SRT) as design criteria for wastewater treatment plants for the removal of endocrine disruptors and pharmaceuticals from wastewater. Water Sci. Technol. 2004, 50, 149–156. [Google Scholar] [CrossRef]
  55. Grandclément, C.; Seyssiecq, I.; Piram, A.; Wong-Wah-Chung, P.; Vanot, G.; Tiliacos, N.; Roche, N.; Doumenq, P. From the conventional biological wastewater treatment to hybrid processes, the evaluation of organic micropollutant removal: A review. Water Res. 2017, 111, 297–317. [Google Scholar] [CrossRef] [Green Version]
  56. Kooijman, G.; de Kreuk, M.K.; Houtman, C.; van Lier, J.B. Perspectives of coagulation/flocculation for the removal of pharmaceuticals from domestic wastewater: A critical view at experimental procedures. J. Water Process Eng. 2020, 34, 101161. [Google Scholar] [CrossRef]
  57. Stackelberg, P.E.; Gibs, J.; Furlong, E.T.; Meyer, M.T.; Zaugg, S.D.; Lippincott, R.L. Efficiency of conventional drinking-water-treatment processes in removal of pharmaceuticals and other organic compounds. Sci. Total. Environ. 2007, 377, 255–272. [Google Scholar] [CrossRef] [PubMed]
  58. Choi, K.-J.; Kim, S.-G.; Kim, S.-H. Removal of antibiotics by coagulation and granular activated carbon filtration. J. Hazard. Mater. 2008, 151, 38–43. [Google Scholar] [CrossRef]
  59. Martínez-Espinosa, R.M. Microorganisms and Their Metabolic Capabilities in the Context of the Biogeochemical Nitrogen Cycle at Extreme Environments. Int. J. Mol. Sci. 2020, 21, 4228. [Google Scholar] [CrossRef] [PubMed]
  60. Molina, M.C.; Bautista, L.F.; Catala, M.; de las Heras, M.R.; Martinez-Hidalgo, P.; San-Sebastian, J.; Gonzalez-Benitez, N. From Laboratory Tests to the Ecoremedial System: The Importance of Microorganisms in the Recovery of PPCPs-Disturbed Ecosystems. Appl. Sci. 2020, 10, 3391. [Google Scholar] [CrossRef]
  61. Nancharaiah, Y.V.; Sarvajith, M.; Mohan, T.V.K. Aerobic Granular Sludge:The Future of Wastewater Treatment. Curr. Sci. 2019, 117, 395. [Google Scholar] [CrossRef]
  62. Muter, O.; Perkons, I.; Selga, T.; Berzins, A.; Gudra, D.; Radovica-Spalvina, I.; Bartkevics, V. Removal of pharmaceuticals from municipal wastewaters at laboratory scale by treatment with activated sludge and biostimulation. Sci. Total Environ. 2017, 584–585, 402–413. [Google Scholar] [CrossRef] [PubMed]
  63. Marchlewicz, A.; Domaradzka, D.; Guzik, U.; Wojcieszyńska, D. Bacillus thuringiensis B1(2015b) is a Gram-Positive Bacteria Able to Degrade Naproxen and Ibuprofen. Water Air Soil Pollut. 2016, 227, 197. [Google Scholar] [CrossRef] [Green Version]
  64. Onesios, K.M.; Yu, J.T.; Bouwer, E.J. Biodegradation and removal of pharmaceuticals and personal care products in treatment systems: A review. Biodegradation 2009, 20, 441–466. [Google Scholar] [CrossRef]
  65. Raza, F.; Zafar, H.; Zhu, Y.; Ren, Y.; -Ullah, A.; Khan, A.U.; He, X.; Han, H.; Aquib, M.; Boakye-Yiadom, K.O.; et al. A Review on Recent Advances in Stabilizing Peptides/Proteins upon Fabrication in Hydrogels from Biodegradable Polymers. Pharmaceutics 2018, 10, 16. [Google Scholar] [CrossRef] [Green Version]
  66. George, A.; Shah, P.A.; Shrivastav, P.S. Natural biodegradable polymers based nano-formulations for drug delivery: A review. Int. J. Pharm. 2019, 561, 244–264. [Google Scholar] [CrossRef] [PubMed]
  67. Neale, R.E.; Barnes, P.W.; Robson, T.M.; Neale, P.J.; Williamson, C.E.; Zepp, R.G.; Wilson, S.R.; Madronich, S.; Andrady, A.L.; Heikkilä, A.M.; et al. Environmental effects of stratospheric ozone depletion, UV radiation, and interactions with climate change: UNEP Environmental Effects Assessment Panel, Update 2020. Photochem. Photobiol. Sci. 2021, 20, 1–67. [Google Scholar] [CrossRef] [PubMed]
  68. Yasri, N.; Hu, J.; Kibria, M.G.; Roberts, E.P.L. Electrocoagulation Separation Processes. In ACS Symposium Series; Chernyshova, I., Ponnurangam, S., Liu, Q., Eds.; American Chemical Society: Washington, DC, USA, 2020; Volume 1348, pp. 167–203. [Google Scholar] [CrossRef]
  69. Aitbara, A.; Khelalfa, A.; Bendaia, M.; Abrane, R.; Amrane, A.; Hazourli, S. Treatment of dairy wastewater by electrocoagulation using A-U4G (2017-Al) alloy and pure aluminum as electrode material. Euro-Mediterr. J. Environ. Integr. 2021, 6, 19. [Google Scholar] [CrossRef]
  70. Dey, T.K.; Uddin, M.E.; Jamal, M. Detection and removal of microplastics in wastewater: Evolution and impact. Environ. Sci. Pollut. Res. 2021, 28, 16925–16947. [Google Scholar] [CrossRef]
  71. Ensano, B.; Borea, L.; Naddeo, V.; Belgiorno, V.; de Luna, M.; Ballesteros, F. Removal of Pharmaceuticals from Wastewater by Intermittent Electrocoagulation. Water 2017, 9, 85. [Google Scholar] [CrossRef] [Green Version]
  72. Nariyan, E.; Aghababaei, A.; Sillanpää, M. Removal of pharmaceutical from water with an electrocoagulation process; effect of various parameters and studies of isotherm and kinetic. Sep. Purif. Technol. 2017, 188, 266–281. [Google Scholar] [CrossRef]
  73. Yoosefian, M.; Ahmadzadeh, S.; Aghasi, M.; Dolatabadi, M. Optimization of electrocoagulation process for efficient removal of ciprofloxacin antibiotic using iron electrode; kinetic and isotherm studies of adsorption. J. Mol. Liq. 2017, 225, 544–553. [Google Scholar] [CrossRef]
  74. Nageeb, M. Adsorption Technique for the Removal of Organic Pollutants from Water and Wastewater. In Organic Pollutants-Monitoring, Risk and Treatment; Rashed, M.N., Ed.; InTech: West Palm Beach, FL, USA, 2013. [Google Scholar] [CrossRef] [Green Version]
  75. Önal, Y.; Akmil-Başar, C.; Sarıcı-Özdemir, Ç. Elucidation of the naproxen sodium adsorption onto activated carbon prepared from waste apricot: Kinetic, equilibrium and thermodynamic characterization. J. Hazard. Mater. 2007, 148, 727–734. [Google Scholar] [CrossRef] [PubMed]
  76. Mondal, S.; Patel, S.; Majumder, S.K. Bio-extract assisted in-situ green synthesis of Ag-RGO nanocomposite film for enhanced naproxen removal. Korean J. Chem. Eng. 2020, 37, 274–289. [Google Scholar] [CrossRef]
  77. Husein, D.Z.; Hassanien, R.; Al-Hakkani, M.F. Green-synthesized copper nano-adsorbent for the removal of pharmaceutical pollutants from real wastewater samples. Heliyon 2019, 5, e02339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Guedidi, H.; Reinert, L.; Soneda, Y.; Bellakhal, N.; Duclaux, L. Adsorption of ibuprofen from aqueous solution on chemically surface-modified activated carbon cloths. Arab. J. Chem. 2017, 10, S3584–S3594. [Google Scholar] [CrossRef] [Green Version]
  79. Baccar, R.; Sarrà, M.; Bouzid, J.; Feki, M.; Blánquez, P. Removal of pharmaceutical compounds by activated carbon prepared from agricultural by-product. Chem. Eng. J. 2012, 211–212, 310–317. [Google Scholar] [CrossRef]
  80. Jafari, K.; Heidari, M.; Rahmanian, O. Wastewater treatment for Amoxicillin removal using magnetic adsorbent synthesized by ultrasound process. Ultrason. Sonochemistry 2018, 45, 248–256. [Google Scholar] [CrossRef]
  81. Hasan, Z.; Choi, E.-J.; Jhung, S.H. Adsorption of naproxen and clofibric acid over a metal–organic framework MIL-101 functionalized with acidic and basic groups. Chem. Eng. J. 2013, 219, 537–544. [Google Scholar] [CrossRef]
  82. Thiebault, T.; Boussafir, M.; Fougère, L.; Destandau, E.; Monnin, L.; Le Milbeau, C. Clay minerals for the removal of pharmaceuticals: Initial investigations of their adsorption properties in real wastewater effluents. Environ. Nanotechnol. Monit. Manag. 2019, 12, 100266. [Google Scholar] [CrossRef]
  83. Styszko, K.; Nosek, K.; Motak, M.; Bester, K. Preliminary selection of clay minerals for the removal of pharmaceuticals, bisphenol A and triclosan in acidic and neutral aqueous solutions. Comptes Rendus Chim. 2015, 18, 1134–1142. [Google Scholar] [CrossRef]
  84. Dordio, A.V.; Miranda, S.; Ramalho, J.P.P.; Carvalho, A.J.P. Mechanisms of removal of three widespread pharmaceuticals by two clay materials. J. Hazard. Mater. 2017, 323, 575–583. [Google Scholar] [CrossRef] [PubMed]
  85. Peiris, C.; Gunatilake, S.R.; Mlsna, T.E.; Mohan, D.; Vithanage, M. Biochar based removal of antibiotic sulfonamides and tetracyclines in aquatic environments: A critical review. Bioresour. Technol. 2017, 246, 150–159. [Google Scholar] [CrossRef]
  86. Wang, J.; Wang, S. Preparation, modification and environmental application of biochar: A review. J. Clean. Prod. 2019, 227, 1002–1022. [Google Scholar] [CrossRef]
  87. Mondal, S.; Patel, S.; Majumder, S.K. Naproxen Removal Capacity Enhancement by Transforming the Activated Carbon into a Blended Composite Material. Water Air Soil. Pollut. 2020, 231, 37. [Google Scholar] [CrossRef]
  88. Essandoh, M.; Kunwar, B.; Pittman, C.U.; Mohan, D.; Mlsna, T. Sorptive removal of salicylic acid and ibuprofen from aqueous solutions using pine wood fast pyrolysis biochar. Chem. Eng. J. 2015, 265, 219–227. [Google Scholar] [CrossRef]
  89. Wang, J.L.; Xu, L.J. Advanced Oxidation Processes for Wastewater Treatment: Formation of Hydroxyl Radical and Application. Crit. Rev. Environ. Sci. Technol. 2012, 42, 251–325. [Google Scholar] [CrossRef]
  90. Wang, J.; Zhuan, R. Degradation of antibiotics by advanced oxidation processes: An overview. Sci. Total. Environ. 2020, 701, 135023. [Google Scholar] [CrossRef]
  91. Wang, J.; Wang, S. Reactive species in advanced oxidation processes: Formation, identification and reaction mechanism. Chem. Eng. J. 2020, 401, 126158. [Google Scholar] [CrossRef]
  92. Ikehata, K.; Naghashkar, N.J.; El-Din, M.G. Degradation of Aqueous Pharmaceuticals by Ozonation and Advanced Oxidation Processes: A Review. Ozone: Sci. Eng. 2006, 28, 353–414. [Google Scholar] [CrossRef]
  93. Aloulou, W.; Aloulou, H.; Jadda, A.; Chakraborty, S.; Amar, R.B. Characterization of an asymmetric ultrafiltration membrane prepared from TiO2-smectite nanocomposites doped with commercial TiO2 and its application to the treatment of textile wastewater. Euro-Mediterr. J. Environ. Integr. 2020, 5, 10. [Google Scholar] [CrossRef]
  94. Friedmann, D.; Mendive, C.; Bahnemann, D. TiO2 for water treatment: Parameters affecting the kinetics and mechanisms of photocatalysis. Appl. Catal. B Environ. 2010, 99, 398–406. [Google Scholar] [CrossRef]
  95. Kumar, A. A Review on the Factors Affecting the Photocatalytic Degradation of Hazardous Materials. Mater. Sci. Eng. Int. J. 2017, 1, 1–10. [Google Scholar] [CrossRef] [Green Version]
  96. Gómez-Avilés, A.; Peñas-Garzón, M.; Bedia, J.; Dionysiou, D.D.; Rodríguez, J.J.; Belver, C. Mixed Ti-Zr metal-organic-frameworks for the photodegradation of acetaminophen under solar irradiation. Appl. Catal. B Environ. 2019, 253, 253–262. [Google Scholar] [CrossRef]
  97. Trovó, A.G.; Melo, S.A.S.; Nogueira, R.F.P. Photodegradation of the pharmaceuticals amoxicillin, bezafibrate and paracetamol by the photo-Fenton process—Application to sewage treatment plant effluent. J. Photochem. Photobiol. A Chem. 2008, 198, 215–220. [Google Scholar] [CrossRef]
  98. Jallouli, N.; Elghniji, K.; Trabelsi, H.; Ksibi, M. Photocatalytic degradation of paracetamol on TiO2 nanoparticles and TiO2/cellulosic fiber under UV and sunlight irradiation. Arab. J. Chem. 2017, 10, S3640–S3645. [Google Scholar] [CrossRef] [Green Version]
  99. Borràs-Ferrís, J.; Sánchez-Tovar, R.; Blasco-Tamarit, E.; Muñoz-Portero, M.J.; Fernández-Domene, R.M.; García-Antón, J. TiO2 Nanostructures for Photoelectrocatalytic Degradation of Acetaminophen. Nanomaterials 2019, 9, 583. [Google Scholar] [CrossRef] [Green Version]
  100. Jallouli, N.; Pastrana-Martínez, L.M.; Ribeiro, A.R.; Moreira, N.F.F.; Faria, J.L.; Hentati, O.; Silva, A.M.T.; Ksibi, M. Heterogeneous photocatalytic degradation of ibuprofen in ultrapure water, municipal and pharmaceutical industry wastewaters using a TiO2/UV-LED system. Chem. Eng. J. 2018, 334, 976–984. [Google Scholar] [CrossRef]
  101. Monteoliva-García, A.; Martín-Pascual, J.; Muñío, M.M.; Poyatos, J.M. Removal of carbamazepine, ciprofloxacin and ibuprofen in real urban wastewater by using light-driven advanced oxidation processes. Int. J. Environ. Sci. Technol. 2019, 16, 6005–6018. [Google Scholar] [CrossRef]
  102. He, Y.; Sutton, N.B.; Rijnaarts, H.H.H.; Langenhoff, A.A.M. Degradation of pharmaceuticals in wastewater using immobilized TiO2 photocatalysis under simulated solar irradiation. Appl. Catal. B Environ. 2016, 182, 132–141. [Google Scholar] [CrossRef]
  103. Calza, P.; Sakkas, V.; Medana, C.; Baiocchi, C.; Dimou, A.; Pelizzetti, E.; Albanis, T. Photocatalytic degradation study of diclofenac over aqueous TiO2 suspensions. Appl. Catal. B Environ. 2006, 67, 3–4. [Google Scholar] [CrossRef]
  104. Jallouli, N.; Elghniji, K.; Hentati, O.; Ribeiro, A.R.; Silva, A.M.T.; Ksibi, M. UV and solar photo-degradation of naproxen: TiO2 catalyst effect, reaction kinetics, products identification and toxicity assessment. J. Hazard. Mater. 2016, 304, 329–336. [Google Scholar] [CrossRef]
  105. Bautitz, I.R.; Nogueira, R.F.P. Degradation of tetracycline by photo-Fenton process—Solar irradiation and matrix effects. J. Photochem. Photobiol. A Chem. 2007, 187, 33–39. [Google Scholar] [CrossRef]
  106. Chen, Q.; Chen, L.; Qi, J.; Tong, Y.; Lv, Y.; Xu, C.; Ni, J.; Liu, W. Photocatalytic degradation of amoxicillin by carbon quantum dots modified K2Ti6O13 nanotubes: Effect of light wavelength. Chin. Chem. Lett. 2019, 30, 1214–1218. [Google Scholar] [CrossRef]
  107. Elmolla, E.S.; Chaudhuri, M. Degradation of amoxicillin, ampicillin and cloxacillin antibiotics in aqueous solution by the UV/ZnO photocatalytic process. J. Hazard. Mater. 2010, 173, 445–449. [Google Scholar] [CrossRef]
  108. Li, J.; Deng, X.; Guo, R.; Li, B.; Cheng, Q.; Cheng, X. Visible light driven photocatalytic decomposition of penicillin G by Ti3+ self-doped TiO2 nano-catalyst through response surface methodology. J. Taiwan Inst. Chem. Eng. 2018, 87, 174–181. [Google Scholar] [CrossRef]
  109. Silva, A.R.; Martins, P.M.; Teixeira, S.; Carabineiro, S.A.C.; Kuehn, K.; Cuniberti, G.; Alves, M.M.; Pereira, L.R. Ciprofloxacin wastewater treated by UVA photocatalysis: Contribution of irradiated TiO2 and ZnO nanoparticles on the final toxicity as assessed by Vibrio fischeri. RSC Adv. 2016, 6, 95494–95503. [Google Scholar] [CrossRef]
  110. Tran, M.L.; Fu, C.-C.; Juang, R.-S. Removal of metronidazole and amoxicillin mixtures by UV/TiO2 photocatalysis: An insight into degradation pathways and performance improvement. Environ. Sci Pollut Res. 2019, 26, 11846–11855. [Google Scholar] [CrossRef] [PubMed]
  111. Tran, M.L.; Fu, C.-C.; Juang, R.-S. Effects of water matrix components on degradation efficiency and pathways of antibiotic metronidazole by UV/TiO2 photocatalysis. J. Mol. Liq. 2019, 276, 32–38. [Google Scholar] [CrossRef]
  112. Zheng, X.; Xu, S.; Wang, Y.; Sun, X.; Gao, Y.; Gao, B. Enhanced degradation of ciprofloxacin by graphitized mesoporous carbon (GMC)-TiO2 nanocomposite: Strong synergy of adsorption-photocatalysis and antibiotics degradation mechanism. J. Colloid Interface Sci. 2018, 527, 202–213. [Google Scholar] [CrossRef] [PubMed]
  113. Molinari, R.; Pirillo, F.; Loddo, V.; Palmisano, L. Heterogeneous photocatalytic degradation of pharmaceuticals in water by using polycrystalline TiO2 and a nanofiltration membrane reactor. Catal. Today 2006, 118, 205–213. [Google Scholar] [CrossRef] [Green Version]
  114. De la Cruz, N.; Dantas, R.F.; Giménez, J.; Esplugas, S. Photolysis and TiO2 photocatalysis of the pharmaceutical propranolol: Solar and artificial light. Appl. Catal. B Environ. 2013, 130–131, 249–256. [Google Scholar] [CrossRef]
  115. Valari, M.; Antoniadis, A.; Mantzavinos, D.; Poulios, I. Photocatalytic reduction of Cr(VI) over titania suspensions. Catal. Today 2015, 252, 190–194. [Google Scholar] [CrossRef]
  116. Serna-Galvis, E.A.; Giraldo-Aguirre, A.L.; Silva-Agredo, J.; Flórez-Acosta, O.A.; Torres-Palma, R.A. Removal of antibiotic cloxacillin by means of electrochemical oxidation, TiO2 photocatalysis, and photo-Fenton processes: Analysis of degradation pathways and effect of the water matrix on the elimination of antimicrobial activity. Environ. Sci. Pollut. Res. 2017, 24, 6339–6352. [Google Scholar] [CrossRef]
  117. Davididou, K.; Monteagudo, J.M.; Chatzisymeon, E.; Durán, A.; Expósito, A.J. Degradation and mineralization of antipyrine by UV-A LED photo-Fenton reaction intensified by ferrioxalate with addition of persulfate. Sep. Purif. Technol. 2017, 172, 227–235. [Google Scholar] [CrossRef] [Green Version]
  118. Almomani, F.A.; Shawaqfah, M.; Bhosale, R.R.; Kumar, A. Removal of emerging pharmaceuticals from wastewater by ozone-based advanced oxidation processes. Environ. Prog. Sustain. Energy 2016, 35, 982–995. [Google Scholar] [CrossRef]
  119. Dantas, R.F.; Canterino, M.; Marotta, R.; Sans, C.; Esplugas, S.; Andreozzi, R. Bezafibrate removal by means of ozonation: Primary intermediates, kinetics, and toxicity assessment. Water Res. 2007, 41, 2525–2532. [Google Scholar] [CrossRef]
  120. Alum, A.; Yoon, Y.; Westerhoff, P.; Abbaszadegan, M. Oxidation of bisphenol A, 17?-estradiol, and 17?-ethynyl estradiol and byproduct estrogenicity. Environ. Toxicol. 2004, 19, 257–264. [Google Scholar] [CrossRef] [PubMed]
  121. Carballa, M.; Manterola, G.; Larrea, L.; Ternes, T.; Omil, F.; Lema, J.M. Influence of ozone pre-treatment on sludge anaerobic digestion: Removal of pharmaceutical and personal care products. Chemosphere 2007, 67, 1444–1452. [Google Scholar] [CrossRef]
  122. Zhao, Y.; Kuang, J.; Zhang, S.; Li, X.; Wang, B.; Huang, J.; Deng, S.; Wang, Y.; Yu, G. Ozonation of indomethacin: Kinetics, mechanisms and toxicity. J. Hazard. Mater. 2017, 323, 460–470. [Google Scholar] [CrossRef]
  123. Koltsakidou, A.; Antonopoulou, M.; Evgenidou, E.; Konstantinou, I.; Lambropoulou, D.A. Cytarabine degradation by simulated solar assisted photocatalysis using TiO2. Chem. Eng. J. 2017, 316, 823–831. [Google Scholar] [CrossRef]
  124. Wang, J.; Chen, C. Biosorbents for heavy metals removal and their future. Biotechnol. Adv. 2009, 27, 195–226. [Google Scholar] [CrossRef] [PubMed]
  125. Li, D.; Cheng, X.; Yu, X.; Xing, Z. Preparation and characterization of TiO2-based nanosheets for photocatalytic degradation of acetylsalicylic acid: Influence of calcination temperature. Chem. Eng. J. 2015, 279, 994–1003. [Google Scholar] [CrossRef]
  126. Biancullo, F.; Moreira, N.F.F.; Ribeiro, A.R.; Manaia, C.M.; Faria, J.L.; Nunes, O.C.; Castro-Silva, S.M.; Silva, A.M.T. Heterogeneous photocatalysis using UVA-LEDs for the removal of antibiotics and antibiotic resistant bacteria from urban wastewater treatment plant effluents. Chem. Eng. J. 2019, 367, 304–313. [Google Scholar] [CrossRef]
  127. Kitsiou, V.; Zachariadis, G.A.; Lambropoulou, D.A.; Tsiplakides, D.; Poulios, I. Mineralization of the antineoplastic drug carboplatin by heterogeneous photocatalysis with simultaneous synthesis of platinum-modified TiO2 catalysts. J. Environ. Chem. Eng. 2018, 6, 2409–2416. [Google Scholar] [CrossRef]
  128. Ofiarska, A.; Pieczyńska, A.; Borzyszkowska, A.F.; Stepnowski, P.; Siedlecka, E.M. Pt–TiO2-assisted photocatalytic degradation of the cytostatic drugs ifosfamide and cyclophosphamide under artificial sunlight. Chem. Eng. J. 2016, 285, 417–427. [Google Scholar] [CrossRef]
  129. Calza, P.; Medana, C.; Sarro, M.; Rosato, V.; Aigotti, R.; Baiocchi, C.; Minero, C. Photocatalytic degradation of selected anticancer drugs and identification of their transformation products in water by liquid chromatography–high resolution mass spectrometry. J. Chromatogr. A 2014, 1362, 135–144. [Google Scholar] [CrossRef]
  130. Sarkar, S.; Chakraborty, S.; Bhattacharjee, C. Photocatalytic degradation of pharmaceutical wastes by alginate supported TiO2 nanoparticles in packed bed photo reactor (PBPR). Ecotoxicol. Environ. Saf. 2015, 121, 263–270. [Google Scholar] [CrossRef]
  131. Teixeira, S.; Gurke, R.; Eckert, H.; Kühn, K.; Fauler, J.; Cuniberti, G. Photocatalytic degradation of pharmaceuticals present in conventional treated wastewater by nanoparticle suspensions. J. Environ. Chem. Eng. 2016, 4, 287–292. [Google Scholar] [CrossRef]
  132. Giraldo, A.L.; Peñuela, G.A.; Torres-Palma, R.A.; Pino, N.J.; Palominos, R.A.; Mansilla, H.D. Degradation of the antibiotic oxolinic acid by photocatalysis with TiO2 in suspension. Water Res. 2010, 44, 5158–5167. [Google Scholar] [CrossRef] [PubMed]
  133. Koltsakidou, A.; Antonopoulou, M.; Sykiotou, M.; Εvgenidou, Ε.; Konstantinou, I.; Lambropoulou, D.A. Photo-Fenton and Fenton-like processes for the treatment of the antineoplastic drug 5-fluorouracil under simulated solar radiation. Environ. Sci. Pollut. Res. 2017, 24, 4791–4800. [Google Scholar] [CrossRef]
  134. Kıdak, R.; Doğan, Ş. Medium-high frequency ultrasound and ozone based advanced oxidation for amoxicillin removal in water. Ultrason. Sonochemistry 2018, 40, 131–139. [Google Scholar] [CrossRef]
  135. Sundin, A.M.; Linderholm, L.; Hedlund, B.; Joyce, K.B.; Klingspor, K. Advanced Wastewater Treatment for Separation and Removal of Pharmaceutical Residues and Other Hazardous Substances: Needs, Technologies and Impacts: A Government-Commissioned Report; Swedish Environmental Protection Agency (Naturvårdsverket): Stockholm, Sweden, 2017; Report 6803; April 2017; Available online: https://www.naturvardsverket.se/Documents/publikationer6400/978-91-620-6803-5.pdf?pid=21820 (accessed on 15 July 2021).
  136. OECD. Pharmaceutical Residues in Freshwater: Hazards and Policy Responses; OECD: Paris, France, 2019. [Google Scholar] [CrossRef]
  137. Mojiri, A.; Vakili, M.; Farraji, H.; Aziz, S.Q. Combined ozone oxidation process and adsorption methods for the removal of acetaminophen and amoxicillin from aqueous solution; kinetic and optimisation. Environ. Technol. Innov. 2019, 15, 100404. [Google Scholar] [CrossRef]
  138. Sirtori, C.; Zapata, A.; Oller, I.; Gernjak, W.; Agüera, A.; Malato, S. Decontamination industrial pharmaceutical wastewater by combining solar photo-Fenton and biological treatment. Water Res. 2009, 43, 661–668. [Google Scholar] [CrossRef]
  139. Akhtar, J.; Amin, N.S.; Aris, A. Combined adsorption and catalytic ozonation for removal of sulfamethoxazole using Fe2O3/CeO2 loaded activated carbon. Chem. Eng. J. 2011, 170, 136–144. [Google Scholar] [CrossRef]
  140. Sui, Q.; Huang, J.; Lu, S.; Deng, S.; Wang, B.; Zhao, W.; Qiu, Z.; Yu, G. Removal of pharmaceutical and personal care products by sequential ultraviolet and ozonation process in a full-scale wastewater treatment plant. Front. Environ. Sci. Eng. 2014, 8, 62–68. [Google Scholar] [CrossRef]
  141. Salgado, R.; Marques, R.; Noronha, J.P.; Carvalho, G.; Oehmen, A.; Reis, M.A.M. Assessing the removal of pharmaceuticals and personal care products in a full-scale activated sludge plant. Environ. Sci. Pollut. Res. 2012, 19, 1818–1827. [Google Scholar] [CrossRef] [PubMed]
  142. Rivera-Utrilla, J.; Sánchez-Polo, M.; Ferro-García, M.Á.; Prados-Joya, G.; Ocampo-Pérez, R. Pharmaceuticals as emerging contaminants and their removal from water. A review. Chemosphere 2013, 93, 1268–1287. [Google Scholar] [CrossRef] [PubMed]
  143. Illés, E.; Szabó, E.; Takács, E.; Wojnárovits, L.; Dombi, A.; Gajda-Schrantz, K. Ketoprofen removal by O3 and O3/UV processes: Kinetics, transformation products and ecotoxicity. Sci. Total. Environ. 2014, 472, 178–184. [Google Scholar] [CrossRef]
  144. Knopp, G.; Prasse, C.; Ternes, T.A.; Cornel, P. Elimination of micropollutants and transformation products from a wastewater treatment plant effluent through pilot scale ozonation followed by various activated carbon and biological filters. Water Res. 2016, 100, 580–592. [Google Scholar] [CrossRef]
  145. Brienza, M.; Nir, S.; Plantard, G.; Goetz, V.; Chiron, S. Combining micelle-clay sorption to solar photo-Fenton processes for domestic wastewater treatment. Environ. Sci. Pollut. Res. 2019, 26, 18971–18978. [Google Scholar] [CrossRef]
  146. Wang, X.; Bi, W.; Zhai, P.; Wang, X.; Li, H.; Mailhot, G.; Dong, W. Adsorption and photocatalytic degradation of pharmaceuticals by BiOCl x I y nanospheres in aqueous solution. Appl. Surf. Sci. 2016, 360, 240–251. [Google Scholar] [CrossRef]
  147. Rahmani, A.R.; Nematollahi, D.; Samarghandi, M.R.; Samadi, M.T.; Azarian, G. A combined advanced oxidation process: Electrooxidation-ozonation for antibiotic ciprofloxacin removal from aqueous solution. J. Electroanal. Chem. 2018, 808, 82–89. [Google Scholar] [CrossRef]
  148. Lin, L.; Jiang, W.; Bechalany, M.; Nasr, M.; Jarvis, J.; Schaub, T.; Sapkota, R.R.; Miele, P.; Wang, H.; Xu, P. Adsorption and photocatalytic oxidation of ibuprofen using nanocomposites of TiO2 nanofibers combined with BN nanosheets: Degradation products and mechanisms. Chemosphere 2019, 220, 921–929. [Google Scholar] [CrossRef]
  149. Lhotský, O.; Krákorová, E.; Mašín, P.; Žebrák, R.; Linhartová, L.; Křesinová, Z.; Kašlík, J.; Steinová, J.; Rodsand, T.; Filipová, A.; et al. Pharmaceuticals, benzene, toluene and chlorobenzene removal from contaminated groundwater by combined UV/H2O2 photo-oxidation and aeration. Water Res. 2017, 120, 245–255. [Google Scholar] [CrossRef]
  150. Peng, J.; Wu, E.; Wang, N.; Quan, X.; Sun, M.; Hu, Q. Removal of sulfonamide antibiotics from water by adsorption and persulfate oxidation process. J. Mol. Liq. 2019, 274, 632–638. [Google Scholar] [CrossRef]
  151. Prado, M.; Borea, L.; Cesaro, A.; Liu, H.; Naddeo, V.; Belgiorno, V.; Ballesteros, F., Jr. Removal of emerging contaminant and fouling control in membrane bioreactors by combined ozonation and sonolysis. Int. Biodeterior. Biodegrad. 2017, 119, 577–586. [Google Scholar] [CrossRef]
Figure 1. Sources of pharmaceutical contamination.
Figure 1. Sources of pharmaceutical contamination.
Applsci 11 06659 g001
Figure 2. Advantages and disadvantages of different treatment process for the removals of pharmaceuticals.
Figure 2. Advantages and disadvantages of different treatment process for the removals of pharmaceuticals.
Applsci 11 06659 g002
Table 1. Class and chemical structure of the most used pharmaceutical products in Tunisia.
Table 1. Class and chemical structure of the most used pharmaceutical products in Tunisia.
ClassPharmaceuticalsChemical FormulaMolecular Weight g/molUsesStructure
Analgesic/ antipyreticAcetaminophenC8H9NO2151.16Used for mild-to-moderate pain and fever. Applsci 11 06659 i001
AspirinC9H8O4180.16Used in the prevention of arterial and venous thrombosis. Applsci 11 06659 i002
LamotrigineC9H7Cl2N5256.09Used antiseizure medication that is a rare but well-known cause of idiosyncratic liver injury that can be severe and even fatal. Applsci 11 06659 i003
TramadolC16H25NO2263.37Used as a narcotic analgesic for severe pain. Applsci 11 06659 i004
AntibioticAmoxicillinC16H19N3O5S365.4Used in the treatment of mild-to-moderate bacterial infections such as sinusitis, bronchitis, otitis media, cellulitis, and community acquired pneumonia. Applsci 11 06659 i005
AzithromycinC38H72N2O12749Related for the rare instances of acute liver injury. Applsci 11 06659 i006
DoxorubicinC27H29NO4543.5Used in the therapy of several forms of lymphoma, leukemia, sarcoma, and solid organ cancers. Applsci 11 06659 i007
CiprofloxacinC17H18FN3O3331.34Used in the therapy of mild-to-moderate urinary and respiratory tract infections caused by susceptible organisms. Applsci 11 06659 i008
MetronidazoleC6H9N3O3171.15Used in the treatment of many anaerobic and certain protozoan and parasitic infections. Applsci 11 06659 i009
OfloxacinC18H20FN3O4361.4Used in case of rare instances of acute hepatocellular injury Applsci 11 06659 i010
SulfamethoxazoleC10H11N3O3S253.28Used in combination with trimethoprim Applsci 11 06659 i011
TrimethoprimC14H18N4O3290.32Used for mild-to-moderate bacterial infections and as prophylaxis against opportunistic infections Applsci 11 06659 i012
ChloramphenicolC11H12CI2N2O5323.13Now used rarely and reserved for severe, life-threatening infections for which other antibiotics are not available. Applsci 11 06659 i013
Oxolinic acidC13H11NO5261.23Used in urinary tract infections. Applsci 11 06659 i014
SulfamonomethoxineC11H12N4O3S280.31It is a sulfonamide and a member of benzenes. Applsci 11 06659 i015
NSAIDsKetoprofenC16H14O3254.28Used in the treatment of acute pain and chronic arthritis. Applsci 11 06659 i016
NaproxenC14H14O3230.26Used for the reduction of pain, fever, inflammation, and stiffness caused by conditions such as osteoarthritis, kidney stones, rheumatoid arthritis, psoriatic arthritis, gout, ankylosing spondylitis, menstrual cramps, tendinitis, bursitis, and for the treatment of primary dysmenorrhea. Applsci 11 06659 i017
IbuprofenC13H18O2 206.28Treatment of rheumatism and arthritis. Applsci 11 06659 i018
DiclofenacC14H11CI2NO2296.1Used for the therapy of chronic forms of arthritis and mild-to-moderate acute pain. Applsci 11 06659 i019
Mefenamic acidC15H15NO2241.28Used in case of rare instances of clinically apparent, acute liver injury. Applsci 11 06659 i020
EtofinamateC18H18F3NO4369.3Used to treat muscle and joint paint. Applsci 11 06659 i021
CaffeineC8H10N4O2194.19Caffeine is the most widely consumed psychostimulant drug in the world and is mostly consumed in the form of coffee. Applsci 11 06659 i022
IndomethacinC19H16ClNO4357.8Used for chronic inflammatory arthritis. Applsci 11 06659 i023
β_blockersPropranololC16H21NO2259.34Used for the therapy of hypertension, cardiac arrhythmias, angina pectoris, and hyperthyroidism. Applsci 11 06659 i024
BisprololC17H29NO4311.4Used to counteract the effect of bradycardia through reduced reflection wave. (b) Applsci 11 06659 i025
CeliprololC20H33N3O4379.5Used for the management of mild to moderate hypertension and effort-induced angina pectoris. Applsci 11 06659 i026
MetoprololC15H25NO3267.36Used in the treatment of several diseases of the cardiovascular system. Applsci 11 06659 i027
TalinololC20H33N3O3363.5It has been investigated for the basic science of gastrointestinal motility disorder. Applsci 11 06659 i028
Tricyclic antidepressants TCACarbamazepineC15H12N2O236.27Used in therapy of epilepsy and trigeminal neuralgia. Applsci 11 06659 i029
VenlafaxineC17H27NO2277.4Can be associated with transient asymptomatic elevations in serum aminotransferase levels. Applsci 11 06659 i030
FluoxetineC17H18F3NO309.33Used to treat major depressive disorder (MDD), moderate-to-severe bulimia nervosa, obsessive–compulsive disorder (OCD), premenstrual dysphoric disorder (PMDD), panic disorder with or without agoraphobia, and, in combination with olanzapine, for treatment-resistant or bipolar I depression. Applsci 11 06659 i031
SulpirideC15H23N3O4S341.4Used therapeutically as an antidepressant, antipsychotic, and as a digestive aid. Applsci 11 06659 i032
AntineoplasticCarboplatinC6H12O4Pt371.25Used as a chemotherapeutic agent for the treatment of various cancers, mainly ovarian, head and neck and lung cancers. Applsci 11 06659 i033
IfosfamideC7H15Cl2N2O2P261.08It is associated with minor transient serum enzyme elevations and has been linked to cases of acute liver injury, including acute cholestatic hepatitis and veno-occlusive disease. Applsci 11 06659 i034
CyclophosphamideC7H15Cl2N2O2P261.08It is associated with minor transient serum enzyme elevations and has been linked to rare cases of acute liver injury. Applsci 11 06659 i035
CytarabineC9H13N3O5243.22Used largely in the therapy of acute leukemia. Applsci 11 06659 i036
MethotrexateC20H22N8O5454.4It has been associated with frequent but mild elevations in serum liver enzymes and, more importantly, with development of chronic liver injury, progressive fibrosis, cirrhosis, and portal hypertension. Applsci 11 06659 i037
Digluconate (chlorhexidine)C34H54Cl2N10O14897.8It is used in various applications, including wound care, hand washing, preoperative body shower, oral hygiene, and general disinfection. Applsci 11 06659 i038
AntisepticChlorhexidineC22H30CI2N10505.4Used as a topical antiseptic and in dental practice for the treatment of inflammatory dental conditions caused by microorganisms. Applsci 11 06659 i039
Anti-convulsantClorazepateC16H11CIN2O3314.72Used as an anticonvulsant as adjunctive therapy in management of epilepsy and as an anxiolytic for therapy of anxiety and alcohol withdrawal. Applsci 11 06659 i040
HydroxyzineC21H27CIN2O2374.9Used largely for symptoms of itching, nausea, anxiety, and tension. Applsci 11 06659 i041
IndapamideC16H16CINO3365.8Antihypertensive agent and a diuretic. Applsci 11 06659 i042
EnalaprilC20H28N2O5376.4Used in the therapy of hypertension and heart failure. Applsci 11 06659 i043
CaptoprilC9H15NO3S217.29Used in the therapy of hypertension and heart failure. Applsci 11 06659 i044
AtenololC14H22N2O3266.34Used as an antihypertensive, hypotensive, and antiarrhythmic, Atenolol acts as a peripheral, cardioselective beta-blocker specific for beta-1 adrenergic receptors, without intrinsic sympathomimetic effects. Applsci 11 06659 i045
BezafibrateC19H20CINO4361.8Used for the treatment of hyperlipidemia. Applsci 11 06659 i046
GabapentinC9H17NO2171.24Used as adjunctive therapy in the management of epilepsy and for neuropathic pain syndromes. Applsci 11 06659 i047
OxcarbazepineC15H12N2O2252.27It is a potent anticonvulsant used alone or in combination with other agents in the therapy of partial seizures. Applsci 11 06659 i048
AntihypertensiveCandesartanC24H20N6O3440.5Used widely in the therapy of hypertension and heart failure. Applsci 11 06659 i049
EprosartanC23H24N2O4S424.5Used for the treatment of high blood pressure. Applsci 11 06659 i050
IrbesartanC25H28N6O428.5It has been linked to rare instances of acute liver injury. Applsci 11 06659 i051
ValsartanC24H29N5O3435.5Used alone or in combination with other agents to treat hypertension and reduce cardiovascular mortality after myocardial infarction. Applsci 11 06659 i052
Data adopted from pubchem (b) [28].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mansouri, F.; Chouchene, K.; Roche, N.; Ksibi, M. Removal of Pharmaceuticals from Water by Adsorption and Advanced Oxidation Processes: State of the Art and Trends. Appl. Sci. 2021, 11, 6659. https://doi.org/10.3390/app11146659

AMA Style

Mansouri F, Chouchene K, Roche N, Ksibi M. Removal of Pharmaceuticals from Water by Adsorption and Advanced Oxidation Processes: State of the Art and Trends. Applied Sciences. 2021; 11(14):6659. https://doi.org/10.3390/app11146659

Chicago/Turabian Style

Mansouri, Fatma, Khawla Chouchene, Nicolas Roche, and Mohamed Ksibi. 2021. "Removal of Pharmaceuticals from Water by Adsorption and Advanced Oxidation Processes: State of the Art and Trends" Applied Sciences 11, no. 14: 6659. https://doi.org/10.3390/app11146659

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop